Titanium Hydrazides Supported by Diamide-Amine ... - ACS Publications

Nov 14, 2008 - this way, four new five-coordinate diamide-amine complexes ... hydrazides containing dianionic N4- or O2N2-donor ligands were also ...
3 downloads 0 Views 509KB Size
Organometallics 2008, 27, 6479–6494

6479

Titanium Hydrazides Supported by Diamide-Amine and Related Ligands: A Combined Experimental and DFT Study Jonathan D. Selby,† Catherine D. Manley,† Andrew D. Schwarz,† Eric Clot,*,‡ and Philip Mountford*,† Chemistry Research Laboratory, UniVersity of Oxford, Mansfield Road, Oxford OX1 3TA, U.K., and Institut Charles Gerhardt Montpellier, UMR 5253 CNRS-UM2-UM1-ENSCM, cc 1501, Place Euge`ne Bataillon, 34095 Montpellier Cedex 5, France ReceiVed August 25, 2008

This paper reports a general method for the synthesis of new terminal titanium diphenyl hydrazido(2-) complexes containing dianionic N3- and N4-donor ligands, along with new hydrazido synthons. Reaction of Ti(NtBu)Cl2(py)3 or Ti(NtBu)Cl2(py′)3 (py′ ) 4-NC5H4tBu) with Ph2NNH2 gave excellent yields of the corresponding monomeric hydrazides Ti(NNPh2)Cl2(L)3 (L ) py (7) or py′), which have been structurally characterized. Application of a dynamic vacuum to 7 formed [Ti(NNPh2)Cl2(py)2]2 (4). Both 4 and 7 are entry points to new titanium hydradizo complexes on reaction with metalated reagents. In this way, four new five-coordinate diamide-amine complexes Ti(NNPh2)(“N2N”)(py) were made (“N2N” Me3SiN(CH2CH2NSiMe3)2, MeN(CH2CH2CH2NSiMe3)2, (2) MeN(CH2CH2NSiMe3)2, NC5H4)C(Me)(CH2NSiMe3)2) and structurally characterized. Five- and six-coordinate terminal titanium hydrazides containing dianionic N4- or O2N2-donor ligands were also synthesized from 4 by an analogous method. The identity of the “N2N” ligand affects the TidNR and NR-Nβ distances of the TidN-NPh2 functional group. A detailed DFT analysis of the bonding in these and a range of model complexes is presented using molecular orbital and natural bond orbital methods. The competition between N(amide) and N(hydrazide) Ti(3dπ)-N(2pπ) interactions has an indirect and significant effect on the NR-Nβ bond. Introduction Group 4 imido complexes (L)MdNR (R is usually alkyl or aryl, L is a supporting ligand (set)) have been a focus of ongoing activity for nearly 20 years.1-6 As well as being of interest from the point of view of fundamental bonding and reactivity (typically focusing on the MdNR bond), practical applications have been found in the catalytic hydroamination of C-C multiple bonds, olefin polymerization, and metal organic chemical vapor deposition. In contrast, the corresponding chemistry of terminal hydrazide complexes (L)MdNNR2 (R is alkyl or aryl) has barely been explored for these early metals. As summarized below, group 4 hydrazide chemistry has nonetheless already established the potential for exciting and significant departures from the norms expected from the imido analogues. However, there are currently no guiding principles for selecting the most appropriate supporting ligand sets to maximize the potential of this chemistry. This paper reports the synthesis and electronic structures of a new class of titanium hydrazide complex for advancing the chemistry of the TidNNR2 functional group. From a more general perspective, the chemistry of metalbound NNR2 species is of considerable interest in the context * Corresponding authors. E-mail: [email protected]; [email protected]. † University of Oxford. ‡ Institut Charles Gerhardt Montpellier. (1) Wigley, D. E. Prog. Inorg. Chem. 1994, 42, 239. (2) Duncan, A. P.; Bergman, R. G. Chem. Rec. 2002, 2, 431. (3) Odom, A. L. Dalton Trans. 2005, 225. (4) Hazari, N.; Mountford, P. Acc. Chem. Res. 2005, 38, 839. (5) Bolton, P. D.; Mountford, P. AdV. Synth. Catal. 2005, 347, 355. (6) Fout, A. R.; Kilgore, U. J.; Mindiola, D. J. Chem.-Eur. J. 2007, 13, 9428.

of the biological and synthetic fixation and activation of dinitrogen.7-15 Chatt, Pickett, and co-workers showed experimentally that the conversion of N2 to NH3 starting from highly reducing zerovalent group 6 tetraphosphine complexes proceeds via (L)M-NNH2 species.16-19 Using a tris(amide)-amine ligand platform, Schrock et al. demonstrated the catalytic conversion of N2 to NH3, again via a Mo-NNH2 intermediate, employing sequential reduction and protonation steps.10,20 Tuczek has reported detailed DFT studies of both the Chatt21 and Schrock22 catalytic cycles, confirming the importance of M-NNH2 species and the reduction-assisted NR-Nβ bond cleavage. The stoichiometric reactions of group 6 M-NNH2 species with organic substrates have also been comprehensively studied. Reactions typically take place at the terminal β-NH2 group, via condensation reactions or insertion into N-H.13-15 Subsequent reductive NR-Nβ bond cleavage can yield organic fragments such as (7) Fryzuk, M. D. Chem. Rec. 2003, 3, 2. (8) Kozak, C. M.; Mountford, P. Angew. Chem., Int. Ed. 2004, 43, 1186. (9) MacKay, B. A.; Fryzuk, M. D. Chem. ReV. 2004, 104, 385. (10) Schrock, R. R. Acc. Chem. Res. 2005, 38, 955. (11) Chirik, P. J. Dalton Trans. 2007, 16. (12) Ohki, Y.; Fryzuk, M. D. Angew. Chem., Int. Ed. 2007, 46, 3180. (13) Fryzuk, M. D.; Johnson, S. A. Coord. Chem. ReV. 2000, 200202, 379. (14) Hidai, M. Chem. ReV. 1995, 95, 1115. (15) Hidai, M. Coord. Chem. ReV. 1999, 185-186, 99. (16) Chatt, J.; Pearman, A. J.; Richards, R. L. Nature 1975, 253, 39. (17) Pickett, C. J.; Leigh, G. J. J. Chem. Soc., Chem. Commun. 1981, 1033. (18) Hussain, W.; Leigh, G. J.; Pickett, C. J. J. Chem. Soc., Chem. Commun. 1982, 747. (19) Pickett, C. J.; Talamarin, J. Nature 1985, 317, 652. (20) Yandulov, D. V.; Schrock, R. R. Science 2003, 76. (21) Stephan, G. C.; Sivasankar, C.; Studt, F.; Tuczek, F. Chem.-Eur. J. 2008, 14, 644. (22) Studt, F.; Tuczek, F. Angew. Chem., Int. Ed. 2005, 44, 5639.

10.1021/om8007597 CCC: $40.75  2008 American Chemical Society Publication on Web 11/14/2008

6480 Organometallics, Vol. 27, No. 24, 2008

pyrroles and pyridines. Notably, direct reaction of the M-NR bond with unsaturated substrates is not seen. Furthermore, structural,23,24 spectroscopic,25 and DFT computational21,22,26,27 data all suggest that the NNR2 ligand in these group 6 complexes is best viewed as a neutral isodiazene ligand (:NdNR2) rather than a hydrazide(2-) species [N-NR2]2-. The first reported terminal titanium hydrazide was Cp2Ti{NN(SiMe3)2}, prepared from Cp2TiCl2 and the thermally sensitive N2(SiMe3)2 (dec > -35 °C).28 There have been no reactivity or structural reports for this system, and the inconvenient nature of N2(SiMe3)2 may have limited its development. Bergman reported and structurally characterized the first zirconium hydrazide, namely, Cp2Zr(NNPh2)(DMAP),29 and Odom was the first to report a structurally authenticated titanium hydrazide.30 Since this report in 2004, only a handful of other terminalgroup4hydrazideshavebeenstructurallycharacterized.31-35 As part of this early work, we reported structural and DFT studies of homologous TidNR (imido) and TidNNPh2 systems, the first study of this type for group 4. These showed that for the early transition metals, NNPh2 is best viewed as a hydrazide(2-) ligand analogous to imide(2-), rather than an isodiazene moiety.31 One could expect early and later metal-NNR2 systems (i.e., hydrazide and isodiazene) to differ in much the same way as Schrock alkylidenes and Fischer carbenes.36 Because of its apparent relationship to the corresponding TidNR systems, one of the main interests in the group 4 MdNNR2 functional group concerns its reactivity with unsaturated organic substrates.37 Bergman’s transient Cp2Zr(NNPh2) system underwent NR-Nβ bond cleavage reactions with alkynes and CO. Surprisingly, ZrdNNR2/RCCR cycloaddition products (well known2 for the corresponding imido systems Cp2Zr(NR)) were not observed. Subsequent work from our own group used the macrocycle-supported complexes Ti(NNPh2)(Me4taa) (H2Me4taa ) tetramethyldibenzotetraaza[14]annulene38). On reaction with TolNCO or CO2, these gave the first well-defined MdNR [2+2] cycloaddition reactions of any group 4 MdNNR2 functional group.39 Woo et al. reported hydrazide ligand transfer reactions for the corresponding Ti(NNR2)(TTP) (H2TTP ) tetrap-tolylporphyrin) system with p-chlorobenzaldehyde, which gave the corresponding hydrazones, also with the NR-Nβ bond remaining intact.40 In contrast, Gade and co-workers once again encountered hydrazide NR-Nβ bond cleavage reactions on reaction of the diamide-amine-supported Zr(N2Npy*) (NNPh2)(py) (1, eq 1) with tBuNC, Ph3PSe, or propylene sulfide (N2Npy* ) (2-NC5H4)C(Me)(CH2NSiMe2tBu)2).35 (23) Allen, F. H.; Kennard, O. Chem. Des. Automation News 1993, 8, 1. and 31. (24) Fletcher, D. A.; McMeeking, R. F.; Parkin, D. J. Chem. Inf. Comput. Sci. 1996, 36, 746 (The United Kingdom Chemical Database Service). (25) Lehnert, N.; Tuczek, F. Inorg. Chem. 1999, 38, 1659. (26) Mersmann, K.; Horn, K. H.; Bores, N.; Lehnert, N.; Studt, F.; Paulat, F.; Peters, G.; Ivanovic-Burmazovic, I.; van Eldik, R.; Tuczek, F. Inorg. Chem. 2005, 44, 3031. (27) Lehnert, N.; Tuczek, F. Inorg. Chem. 1999, 38, 1671. (28) Wiberg, N.; Haring, H.-W.; Huttner, G.; Friedrich, P. Chem. Ber. 1978, 111, 2708. (29) Walsh, P. J.; Carney, M. J.; Bergman, R. G. J. Am. Chem. Soc. 1991, 113, 6343. (30) Li, Y.; Shi, Y.; Odom, A. L. J. Am. Chem. Soc. 2004, 126, 1794. (31) Parsons, T. B.; Hazari, N.; Cowley, A. R.; Green, J. C.; Mountford, P. Inorg. Chem. 2005, 44, 8442. (32) Banerjee, S.; Odom, A. L. Organometallics 2006, 25, 3099. (33) Patel, S.; Li, Y.; Odom, A. L. Inorg. Chem. 2007, 46, 6373. (34) Selby, J. D.; Manley, C. D.; Felix, M.; Schwarz, A. D.; Clot, E.; Mountford, P. Chem. Commun. 2007, 4937. (35) Herrmann, H.; Fillol, J. L.; Wadepohl, H.; Gade, L. H. Angew. Chem., Int. Ed. 2007, 46, 8426.

Selby et al.

In addition to the above stoichiometric reactions of isolated complexes, terminal titanium hydrazides have been implicated (butnotdirectlyobserved)inthecatalytichydrohydrazination30,41-47 and iminohydrazination32,46 of alkynes and in the synthesis of indoles3,43,45 and tryptamines.44 In these reactions, typically leading to a hydrazone or its equivalent, a [2+2] cycloaddition reaction between TidNNR2 and an alkyne (or allene) is considered to be the key N-C bond forming step. Little, if anything, is directly known about these processes or the intermediates involved. We recently reported34 that reaction of the diamide-aminesupported complex Ti(N2NC2,Me)(NNPh2)(py) (2, Vide infra) with PhCCMe did not give the expected [2+2] cycloaddition product, and once again a NR-Nβ bond cleavage product, Ti(N2NC2,Me){NC(Ph)C(Me)NPh2}(py) (3), was formed (eq 2). In addition, reaction of 3 with Ph2NNH2 regenerated 2 to form the cis-1,2-diamino alkene PhC(NH2)C(Me)NPh2. This is a unique and unprecedented transformation of alkynes, which was also found to be catalytic in either 2 or 3.

The rate and selectivity of the new transformation in eq 2 (and Bergman and Gade’s earlier, noncatalytic analogues) should undoubtedly depend on the nature of the strongly π-donating diamide-donor (or related) ligand set. Indeed such chelating anionic nitrogen ligands have been very fruitful for developing the chemistry of early transition metal-ligand multiple bond chemistry during the last 10-15 years.38,48-54 We note in particular Schrock’s tris(amide)-amine (“tren”) systems, which were crucial in that group’s discovery of the synthetic catalytic conversion of N2 to NH3.10,49 Polyamide-based ligands have also been important in the growing field of group 4 N2 activation chemistry.12 Therefore, as a platform for future developments in this field, we report in detail here the synthesis and molecular and electronic structures of a range of new titanium hydrazide complexes with N3-, N4-diamide-donor (and related) ligands, along with new entry synthons for titanium hydrazide chemistry. Part of this work has been communicated.34

Results and Discussion Ligand Choices. Chart 1 illustrates the ligands used in this study. These chelating, dianionic ligands should provide welldefined, relatively nonlabile environments for studying compounds of the type (L)TidNNR2, and the hard N- and O-donors are compatible with the desired Ti(+4) oxidation state. They have been widely used in early transition metal chemistry and metal-ligand multiple bond chemistry4,23,24,38,39,50,51,55-65 and (36) Crabtree, R. H. The Organometallic Chemistry of the Transition Metals, 4th ed.; WileyBlackwell: London, 2005. (37) Mindiola, D. J. Angew. Chem., Int. Ed. 2008, 47, 1557.

Titanium Hydrazides

Organometallics, Vol. 27, No. 24, 2008 6481

Chart 1. The Ligands Used in This Contribution and Their Abbreviations (Formal Anionic Charges Omitted for Clarity)

allow variation of the metal’s coordination number and steric accessibility. They differ in bite angle, hemilability (or otherwise), and π-donor ability. Development of a Synthetic Strategy to New Families of Hydrazide Complexes. While hydrazide complexes (L)TidNNR2 are rather limited in number, several different synthetic strategies have been used for their preparation. (38) Mountford, P. Chem. Soc. ReV. 1998, 27, 105. (39) Blake, A. J.; McInnes, J. M.; Mountford, P.; Nikonov, G. I.; Swallow, D.; Watkin, D. J. J. Chem. Soc., Dalton Trans. 1999, 379. (40) Thorman, J. L.; Woo, L. K. Inorg. Chem. 2000, 39, 1301. (41) Johnson, J. S.; Bergman, R. G. J. Am. Chem. Soc. 2001, 123, 2923. (42) Cao, C.; Shi, Y.; Odom, A. L. Org. Lett. 2002, 4, 2853. (43) Ackermann, L.; Born, R. Tetrahedron Lett. 2004, 45, 9541. (44) Khedkar, V.; Tillack, A.; Michalik, M.; Beller, M. Tetrahedron Lett. 2004, 45, 3123. (45) Tillack, A.; Jiao, H.; Garcia Castro, I.; Hartung, C. G.; Beller, M. Chem.-Eur. J. 2004, 10, 2410. (46) Banerjee, S.; Shi, Y.; Cao, C.; Odom, A. L. J. Organomet. Chem. 2005, 690, 5066. (47) Banerjee, S.; Barnea, E.; Odom, A. L. Organometallics 2008, 27, 1005. (48) Brand, H.; Arnold, J. Coord. Chem. ReV. 1995, 140, 137. (49) Schrock, R. R. Acc. Chem. Res. 1997, 30, 9. (50) Gade, L. H. Chem. Commun. 2000, 173. (51) Gade, L. H.; Mountford, P. Coord. Chem. ReV. 2001, 216-217, 65. (52) Bourget-Merle, L.; Lappert, M. F.; Severn, J. R. Chem. ReV. 2002, 102, 3031. (53) Bigmore, H. R.; Lawrence, S. C.; Mountford, P.; Tredget, C. S. Dalton Trans. 2005, 635. (54) Kempe, R. Angew. Chem., Int. Ed. Engl. 2000, 39, 468. (55) Lowes, T. A.; Ward, B. D.; Whannel, R. A.; Dubberley, S. R.; Mountford, P. Chem. Commun. 2005, 113. (56) Mountford, P.; Swallow, D. J. Chem. Soc., Chem. Commun. 1995, 2357. (57) Blake, A. J.; Mountford, P.; Nikonov, G. I.; Swallow, D. Chem. Commun. 1996, 1835. (58) Swallow, D.; Dunn, S. C.; Nikonov, G. I.; Mountford, P. Abstr. Papers Am. Chem. Soc. 1996, 211, 78. (59) Swallow, D.; McInnes, J. M.; Mountford, P. J. Chem. Soc., Dalton Trans. 1998, 2253. (60) McInnes, J. M.; Swallow, D.; Blake, A. J.; Mountford, P. Inorg. Chem. 1998, 37, 5970. (61) Kisko, J. L.; Hascall, T.; Parkin, G. J. Am. Chem. Soc. 1997, 119, 7609. (62) Amgoune, A.; Thomas, C. M.; Roisnel, T.; Carpentier, J.-F. Chem.Eur. J. 2006, 12, 169. (63) Bonnet, F.; Cowley, A. R.; Mountford, P. Inorg. Chem. 2005, 44, 9046. (64) Suzuki, Y.; Terao, H.; Fujita, T. Bull. Chem. Soc. Jpn. 2003, 76, 1493. (65) Mitani, M.; Saito, J.; Ishii, S.; Nakayama, Y.; Makio, H.; Matsukawa, N.; Matsui, S.; Mohri, J.; Furuyama, R.; Terao, H.; Bando, H. H. T.; Fujita, T. Chem. Rec. 2004, 4, 137.

Wiberg28 and Woo40 employed Me3SiCl or LiCl elimination protocols, respectively, starting from the corresponding dichloride precusors (L)TiCl2 (L ) Cp2 or TTP). We showed that Ti(Me4taa)(NtBu) reacted with Ph2NNH2 to form Ti(Me4taa)(NNPh2) and tBuNH2 via a protonolysis type reaction39 reminiscent of the well-established tert-butyl imide/aniline exchange protocols used to form aryl imides.4,6,66-68 Protonolysis reactions of bis(dimethylamido) compounds (L)Ti(NMe2)2 with several different hydrazines RR′NNH2 (R ) R ) alkyl or aryl; R ) aryl, R′ ) H) have also afforded both terminal and bridging hydrazido derivatives.30,33 Although protonolysis procedures appeared attractive for preparing the target diamide-amine hydrazide complexes (avoiding potentially reducing anionic reagents), they were unsatisfactory due to concomitant protonolysis side-reactions of the supporting ligands. For example, reaction of Ti(N2NC2,SiMe3)(NtBu)(py)69 with Ph2NNH2 in C6D6 gave less than ca. 35% conversion to the target Ti(N2NC2,SiMe3)(NNPh2)(py) (Vide infra). The remainder of the Ti(N2NC2,SiMe3)(NtBu)(py) was consumed to form the protio ligand H2N2NC2,SiMe3 and unknown species. Analogous results were found for certain other tert-butyl imido titanium complexes of the target ligands. Gade and co-workers have encountered similar difficulties starting from Ti(N2Npy)(NMe2)2.70 The amido nitrogens of N3- and N4-donor supporting ligands of the type N2NCn,R, N2Npy, and N2NN′ are therefore prone to protonolysis reactions at a rate competitive with those of the target TidNtBu or Ti-NMe2 leaving groups. This is in contrast to the Ti(Me4taa)(NtBu) system, which underwent imide protonolysis with Ph2NNH2 without loss of H2Me4taa.39,59 The difference may be partially attributable to the charge delocalization within the Me4taa backbone, making the formally anionic ligands less Brønsted basic. Odom’s successful conversion of Ti(dmpa)(NMe2)2 to Ti(dmpa)(NNMe2)(tBubipy) might be explained in a similar way (dmpa ) N,N-di(pyrrolyl-R-methyl)-N-methylamine; tBubipy ) di-tert-butylbipyridine).30 (66) Blake, A. J.; Collier, P. E.; Dunn, S. C.; Li, W.-S.; Mountford, P.; Shishkin, O. V. J. Chem. Soc., Dalton Trans. 1997, 1549. (67) Coles, M. P.; Dalby, C. I.; Gibson, V. C.; Clegg, W.; Elsegood, M. R. J. Polyhedron 1995, 14, 2455. (68) Michelman, R. I.; Bergman, R. G.; Andersen, R. A. Organometallics 1993, 12, 2741. (69) Pugh, S. M.; Clark, H. S. C.; Love, J. B.; Blake, A. J.; Cloke, F. G. N.; Mountford, P. Inorg. Chem. 2000, 39, 2001. (70) Gade, L. H. Personal communication.

6482 Organometallics, Vol. 27, No. 24, 2008

Selby et al.

Scheme 1. Synthesis of Titanium Hydrazides Supported by Tridentate Diamide-Amine Ligandsa

Reagents and conditions: (i) Li2N2NC2,Me, -78 °C, toluene, 65% or Li2N2NC2,SiMe3, -78 °C, toluene, 26%; (ii) Li2N2NC3,Me, -78 °C, toluene, 41%; (iii) Li2N2Npy, -78 °C, toluene, 73%. a

Scheme 2. Synthesis of Titanium Hydrazides Supported by Tetradentate N4- or O2N2-Donor Ligandsa

a

Reagents and conditions: (i) Li2N2NN′, -78 °C, toluene, 31%; (ii) Li2Me4taa, C6D6, 100%; (iii) Na2O2NN′, -78 °C, toluene, 49%.

A wide number of titanium imido compounds4,6 are available from synthons of the type Ti(NR)Cl2(py)x or Ti(NR)Cl2(NHMe2)2 (R ) alkyl or aryl; x ) 2 or 3)66,71,72 by substitution strategies. We therefore considered an analogous approach for making new hydrazide complexes. At first sight, a suitable hydrazide synthon is Ti(NNPh2)Cl2(NHMe2)2 (prepared from Ti(NMe2)2Cl2 and Ph2NNH2).31 Unfortunately, the somewhat acidic NH groups of the NHMe2 ligands lead to undesired side-reactions when the complexes are treated with anionic reagents such as lithiated amides.4 However, the NHMe2 (71) Adams, N.; Arts, H. J.; Bolton, P. D.; Cowell, D.; Dubberley, S. R.; Friederichs, N.; Grant, C.; Kranenburg, M.; Sealey, A. J.; Wang, B.; Wilson, P. J.; Cowley, A. R.; Mountford, P.; Schro¨der, M. Chem. Commun. 2004, 434. (72) Adams, N.; Bigmore, H. R.; Blundell, T. L.; Boyd, C. L.; Dubberley, S. R.; Sealey, A. J.; Cowley, A. R.; Skinner, M. E. G.; Mountford, P. Inorg. Chem. 2005, 44, 2882. (73) Lorber, C.; Choukroun, R.; Donnadieu, B. Inorg. Chem. 2002, 41, 4217.

ligands of Ti(NNPh2)Cl2(NHMe2)2 can be substituted by pyridine to form dimeric [Ti(NNPh2)Cl2(py)2]2 (4), believed to have terminal TidNNPh2 groups and bridging Cl ligands.31 Preliminary experiments showed that this material reacted with metalated salts of the target ligands to form the desired complexes in acceptable to good yields as described below (Schemes 1 and 2).

Although [Ti(NNPh2)Cl2(py)2]2 (4) can be prepared from Ti(NNPh2)Cl2(NHMe2)2, the latter itself requires at least a two-

Titanium Hydrazides

Organometallics, Vol. 27, No. 24, 2008 6483

Figure 1. Displacement ellipsoid plots (25% probability) with H atoms omitted for clarity. (a) Ti(NNPh2)Cl2(py)3 (7), atoms carrying the suffix “A” are related to their counterparts by the symmetry operator -x+1, y, -z+3/2. (b) One of the two crystallographically independent molecules of Ti(NNPh2)Cl2(py′)3 (8), second orientation of one of the py′ tert-butyl groups omitted.

step synthesis, and its conversion to 4 requires heating in a large excess of neat pyridine at 70 °C for 16 h, giving variable yields and purity. Similar difficulties with this type of apparently straightforward reaction have been noted for the imido systems M(NR)Cl2(NHMe2)2 (M ) Ti72 or V73). We therefore developed an alternative route (eq 3) to 4 and well-defined monomeric homologues starting from the tert-butyl imido compounds Ti(NtBu)Cl2(L)3 (L ) py (5) or 4-NC5H4tBu (py′, 6)). These are both readily available in multigram quantities from inexpensive and commerically available starting materials (TiCl4, t BuNH2, and py or py′).66,74 Reaction of 5 with Ph2NNH2 (1 equiv) in benzene at room temperature overnight afforded a precipitate of bright yellow, monomeric Ti(NNPh2)Cl2(py)3 (7) in 89% yield (ca. 5 g scale) and in analytically pure form. The corresponding reaction of 6 gave the homologous compound Ti(NNPh2)Cl2(py′)3 (8) in 74% yield, again as a pure yellow solid. The lower isolated yield for 8 reflects its higher solubility in hydrocarbon solvents. Further exposure of 7 to a dynamic vacuum leads to loss of one coordinated pyridine and eventual quantitative formation of the bis(pyridine) complex 4. Analogous behavior is found for the imido systems Ti(NR)Cl2(py)3 (R ) tBu or aryl)66 and was attributed to the strong labilizing effect of the imido groups on the pyridine trans to the TidNR bond.75 Compound 8 does not easily lose the trans py′ ligand in this way because of the lower volatility of the substituted pyridine and its better donor ability. Both 4 and 7 were subsequently used as synthons for the syntheses described below, and there is no difference in their performance. Compound 8 was not used further in the present syntheses because preliminary studies found the parent pyridine adduct products were already very soluble and prone to be oily in some instances (due to the presence of SiMe3 groups in the diamide complexes). Unfortunately, reaction of Me2NNH2 with 5 did not give terminal hydrazido compounds but only the (74) Dunn, S. C.; Batsanov, A. S.; Mountford, P. J. Chem. Soc., Chem. Commun. 1994, 2007. (75) Kaltsoyannis, N.; Mountford, P. J. Chem. Soc., Dalton Trans. 1999, 781.

Table 1. Selected Bond Distances (Å) and Angles (deg) for Ti(NNPh2)Cl2(py)3 (7)a Ti(1)-N(1) Ti(1)-N(3) N(1)-N(2) N(1)-Ti(1)-N(3) N(1)-Ti(1)-Cl(1) N(1)-Ti(1)-N(4) N(3)-Ti(1)-Cl(1)

1.727(2) 2.2464(17) 1.359(3) 97.25(4) 96.738(15) 180 90.44(5)

Ti(1)-Cl(1) Ti(1)-N(4) N(2)-C(1) Cl(1)-Ti(1)-Cl(1A) Ti(1)-N(1)-N(2) N(1)-N(2)-C(1) C(1)-N(2)-C(1A)

2.4265(5) 2.402(2) 1.411(2) 166.52(3) 180 118.66(11) 118.66(11)

a Atoms carrying the suffix “A” are related to their counterparts by the symmetry operator -x+1, y, -z+3/2.

hydrazide-bridged dimer Ti2(µ-η2,η1-NNMe2)2Cl4(py)2, which we found to be ineffective as a synthon.31 The NMR spectra of 7 and 8 show resonances for a NNPh2 ligand and two py (or py′) ligand environments in a 2:1 ratio. The lower intensity py or py′ resonances are shifted upfield from their counterparts and are assigned to the ligands trans to the TidNNPh2 functional groups. The solution NMR and other analytical data are fully consistent with the structures proposed in eq 3, but do not uniquely define them as monomeric terminal hydrazides. They have therefore been characterized by X-ray crystallography in both cases. The molecular structures are shown in Figure 1, and key distances and angles are given in Tables 1 and 2. Molecules of 7 lie on crystallographic 2-fold rotation axes passing through the Ti(1)-N(1)-N(2) linkage, while compound 8 contains two crystallographically independent molecules in the asymmetric unit. Some positional disorder of certain py′ tert-butyl groups of 8 was satisfactorily modeled. The structures of 7 and 8 confirm those anticipated in eq 3. These are only the second and third examples of a structurally authenticated TidNNPh2 functional group (two such examples are now known for zirconium).29,35 The other titanium example is Ti(NNPh2){HC(Me2pz)3}Cl2 (9), which also contains an octahedral Ti, but with fac-coordination of the N3-donor set and mutually cis Cl ligands. Since a number of imido analogues Ti(NR)Cl2(py)3 (R ) tBu or aryl) have been structurally characterized,23,24 7 and 8 provide an excellent opportunity for

6484 Organometallics, Vol. 27, No. 24, 2008

Selby et al.

Table 2. Selected Bond Distances (Å) and Angles (deg) for Ti(NNPh2)Cl2(py′)3 (8)a Ti(1)-N(1) Ti(1)-Cl(1) Ti(1)-Cl(2) Ti(1)-N(3) Ti(1)-N(4) N(1)-Ti(1)-N(3) N(1)-Ti(1)-N(4) N(1)-Ti(1)-N(5) N(1)-Ti(1)-Cl(1) N(1)-Ti(1)-Cl(2) Cl(1)-Ti(1)-Cl(2) a

1.7240(18) 2.4079(7) 2.4224(7) 2.3942(19) 2.2417(18) 178.56(8) 94.51(8) 95.97(8) 96.68(6) 99.23(6) 164.07(3)

[1.7216(17)] [2.4132(7)] [2.3967(6)] [2.483(2)] [2.2314(19)] [178.20(8)] [93.71(8)] [93.41(8)] [98.30(6)] [98.23(6)] [163.43(3)]

Ti(1)-N(5) N(1)-N(2) N(2)-C(1) N(2)-C(7) N(4)-Ti(1)-N(5) Ti(1)-N(1)-N(2) N(1)-N(2)-C(1) N(1)-N(2)-C(7) C(1)-N(2)-C(7)

2.2274(18) 1.351(2) 1.414(3) 1.430(3) 169.51(7) 179.54(15) 118.49(17) 118.69(17) 122.82(17)

[2.211(2)] [1.353(2)] [1.423(3)] [1.417(3)] [172.69(7)] [178.38(16)] [117.77(18)] [118.54(18)] [123.45(18)]

Values in brackets are for the other crystallographically independent molecule.

structural comparisons with their better established imido analogues.

The metric data for the structures of 7 and 8 are generally comparable within error. They possess pseudo-octahedral geometries at Ti(1), mutually trans Cl donors, and three py (or py′) ligands arranged in a mer fashion. This is the same as for the imido analogues Ti(NR)Cl2(py)3.23,24 There is a significant bond-lengthening effect of the hydrazido ligand on the py or py′ ligand opposite it, and the magnitude of this trans influence (av ca. 0.20 Å) is comparable to that found in the imido analogues. The Ti-Cl and Ti-py (py′) distances for 7 and 8 are within the ranges found for these imido and other Ti(+4) compounds.23,24 They are thus consistent with NNPh2 acting as a formally dianionic “hydrazide(2-)” ligand (as opposed to being considered as a neutral NdNPh2 “isodiazene”76) in these titanium complexes, as also found in the DFT calculations for 9.31 The parameters associated with TidNNPh2 linkages are similar to those in 9 with comparable TidNR (1.722(2)-1.727(2) Å) and NR-Nβ (1.351(2)-1.359(3) Å) distances (cf. 1.718(2) and 1.369(3) Å, respectively in 9). Likewise, the hydrazide linkages are linear (TidN-NPh2 ) 178-180 °) and the Nβ nitrogen is planar (sum of the angles subtended at N(2) ∼360 ° within error) for both compounds. The Ph rings are rotated by ca. 30-40° out of the plane containing the TidN-N(Cipso)2 atoms, and the Nβ-Cipso distances are approximately equivalent in all cases (1.41-1.42 Å). This suggests the possibility of partial delocalization of electron density from the sp2-hybridized Nβ onto the phenyl rings. The TidNR distances in 7 and 8 are all slightly longer than for the imido complexes Ti(NR)Cl2(py)3: R ) tBu, TidN ) 1.705(3) Å; R ) Ph, 1.714(2) Å; R ) p-Tol, 1.705(4) Å).66 They are, however, well within the general range (ca. 1.68-1.76 Å based on titanium imides) expected for a formal TitNR triple bond (σ2 π4). This triple-bond assignment is also consistent with previous DFT calculations on 9 and also the observed linear TidN-NPh2 linkage (sp-hybridized NR).31 In contrast, the average NR-Nβ distance for transition metal diphenylhydrazido (76) Kahlal, S.; Saillard, J.-Y.; Hamon, J.-R.; Manzur, C.; Carrillo, D. J. Chem. Soc., Dalton Trans. 1998, 1229.

compounds in general is 1.317 Å,23,24 and so in this sense the NR-Nβ distances for 7 and 8 are comparatively long. Synthesis of Titanium Hydrazides Containing Dianionic N3-, N4-, and O2N2-Donor Ligands. As mentioned, both [Ti(NNPh2)Cl2(py)2]2 (4) and Ti(NNPh2)Cl2(py)3 (7) serve as starting materials for the general synthesis of new families of titanium diphenyl hydrazides containing dianionic N3- and N4-donor ligands. The syntheses are summarized in Schemes 1 and 2. Preliminary studies focused on the N2NC2,SiMe3 ligand system, which we have previously used in the tert-butyl imido complex Ti(NtBu)(N2NC2,SiMe3)(py). Reaction of Li2N2NC2,SiMe3 with 4 in benzene at room temperature gave the crude product as a green-brown solid. The 1H NMR spectrum showed the presence of both the desired complex Ti(NNPh2)(N2NC2,SiMe3)(py) (10) and free protio-ligand H2N2NC2,SiMe3 in a ca. 1:2 ratio. The presence of H2N2NC2,SiMe3 is attributed to redox-coupled H atom abstraction side-reactions, as have been found previously in some analogous reactions of this ligand.69 These could be partially alleviated by starting the reaction at -78 °C and adding precooled toluene to a cold solid mixture of Li2N2NC2,SiMe3 and 4, followed by a short reaction time and workup. Analytically pure samples of 10 were obtained in 26% overall yield, the relatively high solubility of 10 in hydrocarbon solvents also contributing to the disappointing yield. The reactions between 4 and Li2N2NC2,Me or Li2N2NC3,Me were carried out in the same way and afforded the homologous complexes Ti(NNPh2)(N2NC23,Me)(py) (2) and C3,Me Ti(NNPh2)(N2N )(py) (11) in 65% and 41% isolated yield. We found that Li2N2Npy reacted smoothly with 4 to form Ti(NNPh2)(N2Npy)(py) (12) in good yield (73%) without noticeable redox complications. This reaction could best be performed by adding a cold solution of the lithiated amide to 4. The NMR and other analytical data for the new compounds 10-12 are fully consistent with the structures proposed in Scheme 1, which have all been confirmed by X-ray crystallography (Vide infra). We were also interested to prepare complexes with dianionic N4-donor ligands (Scheme 2). Thus reaction of Li2N2NN′ with 4 under analogous conditions for 2, 10, and 11 afforded the five-coordinate Ti(NNPh2)(N2NN′) (13). We were not able to grow diffraction-quality crystals of 13, but its NMR data are consistent with the Cs symmetrical, trigonal-bipyramidal structure proposed in Scheme 2. Titanium and zirconium imido analogues of 13 have been reported previously and have comparable structures.77 The chemical shift of the ortho-H of the pyridyl donor in 13 appears at δ 8.98 ppm, consistent with it being coordinated in solution (in free H2N2NN′ this hydrogen

(77) Skinner, M. E. G.; Toupance, T.; Cowhig, D. A.; Tyrrell, B. R.; Mountford, P. Organometallics 2005, 24, 5586.

Titanium Hydrazides

Organometallics, Vol. 27, No. 24, 2008 6485

Figure 2. Displacement ellipsoid plots with H atoms omitted for clarity. (a) Ti(NNPh2)(N2NC2,Me)(py) (2, 20% probability, disorder in coordinated pyridine omitted). (b) Ti(NNPh2)(N2NC2,SiMe3)(py) (10, 30% probability).

appears at δ 8.60 ppm).78 Significantly, 13 is the neutral group 4 analogue of Schrock’s triamide-amine (tren)-supported Mo-NNR2 (R ) alkyl or H) systems.10,20,79 As mentioned, Ti(NNPh2)(Me4taa) (14) was previously prepared by reaction of Ti(NtBu)(Me4taa) with Ph2NNH2.39 As shown in Scheme 2, the reaction between Li2Me4taa and 4 in C6D6 gave quantitative conversion to 14 immediately at room temperature, indicating this is a viable route to macrocyclesupported hydrazides also. The NMR data for 14 were comparable to those of an authentic sample prepared by the original imide/hydrazine exchange route. We have also made a comparison between the N4-donor systems Ti(NNPh2)(L) (L ) N2NN′ (13) or Me4taa (14)) and those of O2NN′ (see Chart 1) as a representative bis(phenolate)diamine ligand. Addition of a cold toluene solution of Na2O2NN′ to [Ti(NNPh2)Cl2(py)2]2 (4) gave the compound Ti(NNPh2)(O2NN′)(py) (15) in 49% yield as summarized in Scheme 2. The NMR data for 15 are consistent with the Cs symmetric structure proposed and confirm the presence of the coordinated pyridine. Corresponding monomeric titanium imido complexes of O2NN′-type ligands have been prepared previously.80 Unfortunately we have not been able to obtain diffraction-quality crystals of 15, and for the remainder of this paper we shall focus on the diamide-amine systems. Attempts to prepare a terminal dimethyl hydrazide analogue of 15 are described in the Supporting Information along with the X-ray structure of dimeric Ti2(µ-η2η1-NNMe2)2(O2NN′)2 (16). Structures of Titanium Hydrazides Containing N3-Donor Ligands. The compounds 2 and 10-12 offer an excellent opportunity to determine systematically the effects of the different diamide-amine ligands on the ground-state structure of this new class of titanium hydrazide complexes, and thereby gain insight into the unique reactivity of these systems. Diffraction-quality crystals of Ti(NNPh2)(N2NC2,Me)(py) (2), Ti(NNPh2)(N2NC2,SiMe3)(py) (10), Ti(NNPh2)(N2NC3,Me)(py) (11), and Ti(NNPh2)(N2Npy)(py) (12) were grown from the appropriate solvents. The molecular structures are shown in (78) Skinner, M. E. G.; Li, Y.; Mountford, P. Inorg. Chem. 2002, 41, 1110. (79) O’Donoghue, M. B.; Davis, W. M.; Schrock, R. R. Inorg. Chem. 1998, 37, 5149. (80) Boyd, C. L.; Toupance, T.; Tyrrell, B. R.; Ward, B. D.; Wilson, C. R.; Cowley, A. R.; Mountford, P. Organometallics 2005, 24, 309.

Figures 2 (2, 10) and 3 (11, 12). Key bond distances and angles are compared in Table 3. All four complexes possess trigonal-bipyramidal titanium centers with the diamide-amine ligand coordinating in a fac manner. The hydrazide N(1) (NR) and amide nitrogens (N(3), N(4)) occupy the equatorial positions of the trigonal bipyramids; the neutral donors, the axial sites. The Ti-N and intraligand distances and angles for the diamide-amine and pyridine donors are all within the expected ranges,23,24 although there are some interesting variations as discussed below. The titanium tert-butyl imido analogues of 10 and 12 have been reported previously,69,81 as have related group 4 and 5 imido complexes Zr(N-2,6C6H3iPr2)(N2Npy)(py) and Ta(NtBu)(N2Npy)Cl.81,82 These have comparable structures with the amide and imide nitrogens in the equatorial plane. In the hydrazide structures, as in the aforementioned imido complexes, the amide nitrogens are trigonal planar (sum of the angles subtended at N(3) and N(4) ) 360° within error) and sp2 hybridized, which is typical for early transition metal complexes.50,54 Before considering the data associated with the TidNNPh2 moieties it is necessary to examine more closely the variations in the Ti(N2NR)(py) fragments in the four complexes. The fivemembered chelate rings formed by the diamide-amine ligands in 2 and 10 lead to larger N(1)-Ti(1)-N(5) angles (av ca. 103°) compared to those for 11 and 12 (six-membered chelate rings, av ca. 94°), thus pulling the apical donor away from the TidNNPh2 linkage. The N(5)-Ti(1)-N(6) angles are correspondingly smaller for 2 and 10 (av ca. 161°) than for 11 and 12 (av ca. 173°). This in turn affects the Ti(1)-N(6) bond distances for the pyridine ligand (shorter in 2 and 10 (av ca. 2.22 Å) than in 11 and 12 (av ca. 2.27 Å). The effect of replacing the apical N-Me of the diamide ligand in 2 with the bulky N-SiMe3 in 10 results in a 0.131(4) Å lengthening of the Ti(1)-N(5) bond. Interestingly, the Ti(1)-N(6) distances for the pyridine ligands do not vary significantly between 2 and 10 (difference 0.011(5) Å). Finally it is important to note that the variations in the diamide-amine ligand framework lead to significant changes in the Ti(1)-N(3) and Ti(1)-N(4) distances to the amide nitrogens. Thus from 2 to 12 the average Ti-Namide (81) Blake, A. J.; Collier, P. E.; Gade, L. H.; Mountford, P.; Lloyd, J.; Pugh, S. M.; Schubart, M.; Skinner, M. E. G.; Tro¨sch, J. M. Inorg. Chem. 2001, 40, 870. (82) Ward, B. D.; Orde, G.; Clot, E.; Cowley, A. R.; Gade, L. H.; Mountford, P. Organometallics 2004, 23, 4444.

6486 Organometallics, Vol. 27, No. 24, 2008

Selby et al.

Figure 3. Displacement ellipsoid plots with H atoms omitted for clarity. (a) Ti(NNPh2)(N2NC3,Me)(py) (11, 20% probability, minor disorder omitted). (b) Ti(NNPh2)(N2Npy)(py) (12, 20% probability). Table 3. Selected Bond Distances (Å) and Bond and Torsion Angles (deg) for Ti(NNPh2)(N2NC2,Me)(py) (2), Ti(NNPh2)(N2NC2,SiMe3)(py) (10), Ti(NNPh2)(N2NC3,Me)(py) (11), and Ti(NNPh2)(N2Npy)(py) (12)a parameter

2

10

11

12

Ti(1)-N(1) Ti(1)-N(3) Ti(1)-N(4) Ti(1)-N(5) Ti(1)-N(6) N(1)-N(2) N(2)-C(1) N(2)-C(7) N(1)-Ti(1)-N(3) N(1)-Ti(1)-N(4) N(3)-Ti(1)-N(4) N(1)-Ti(1)-N(5) N(1)-Ti(1)-N(6) N(5)-Ti(1)-N(6) Ti(1)-N(1)-N(2) N(1)-N(2)-C(1) N(1)-N(2)-C(7) C(1)-N(2)-C(7) av N(1)-Ti(1)N(3)-R1 av N(1)-Ti(1)N(4)-R2

1.733(5) 2.029(5) 2.026(5) 2.224(4) 2.215(5) 1.359(7) 1.438(7) 1.399(7) 113.8(2) 113.7(2) 130.5(2) 102.1(2) 99.3(2) 158.4(2) 177.7(4) 116.1(4) 120.9(5) 119.5(5) 53.0

1.750(2) 2.014(2) 2.003(2) 2.355(2) 2.226(2) 1.362(2) 1.426(2) 1.403(2) 118.83(7) 120.57(7) 120.29(7) 104.01(6) 93.17(7) 162.82(6) 173.22(14) 118.3(2) 120.2(2) 121.4(2) 38.6

1.7506(14) 1.999(2) 2.0020(14) 2.289(2) 2.281(2) 1.364(2) 1.433(2) 1.398(2) 114.89(6) 116.03(6) 128.98(6) 92.91(6) 95.57(6) 171.49(5) 168.75(13) 116.78(14) 121.25(14) 120.66(14) 37.4

1.759(2) 1.974(2) 1.974(2) 2.184(2) 2.259(2) 1.367(2) 1.393(3) 1.434(3) 124.52(8) 128.74(8) 106.64(7) 95.49(7) 90.25(7) 174.26(6) 163.8(2) 121.2(2) 115.9(2) 122.7(2) 21.2

70.3

49.4

21.5

25.2

a “Av N(1)-Ti(1)-N(3,4)-R1,2” refers to the average of the torsion angles N(1)-Ti(1)-N(3,4)-Si or N(1)-Ti(1)-N(3,4)-C for the linkages under consideration.

distance decreases from ca. 2.027 Å to 1.974 Å. At the same time the orientation of the amide nitrogen substituents relative to the equatorial plane changes such that they become more coplanar with the amide nitrogens and hydrazide NR. This can be quantified by the average N(1)-Ti(1)-Namide-R torsion angles (see Table 3), which decrease from 61.6° in 2 to 23.2° in 12. The significance of these variations is evaluated using DFT later. Turning now to the TidNNPh2 moiety, it is seen that in all compounds the Ti(1)-N(1)-N(2) linkage is approximately linear (163.8(2)-177.7(4)°). The β-nitrogen (N(2)) is effectively trigonal planar, as judged by the sums of the angles subtended at this atom (range 356(1)-360(1)°). These features are also consistent with the structures of 7, 8, and 9. Similarly, the N(1)-N(2) distances in 2 and 10-12 (1.359(7)-1.367(2) Å) fall within the range of those found in the dichloride complexes (range 1.351(2)-1.369(3) Å). In contrast, the Ti(1)-N(1)

distances (1.733(5)-1.759(2) Å) are systematically longer than for 7, 8, and 9 (range 1.718(2)-1.727(2) Å). The same situation is found when comparing the TidNtBu distances of Ti(NtBu)(N2NC2,SiMe3)(py)69 and Ti(NtBu)(N2Npy)(py)81 with that of Ti(NtBu)Cl2(py)366 and in the first instance can be attributed to the good donor abilities of diamide-amine ligands. The TidNR distances in 10 and 12 are longer by 0.032(4) and 0.035(3) Å than the TidNtBu distances in the corresponding imido complexes, as expected from our earlier comparison of 7 with Ti(NtBu)Cl2(py)3. The Nβ-Cipso distances (N(2)-C(1) and N(2)-C(7)) merit comment. In 10 the difference between them is relatively small (0.023(3) Å), and the phenyl groups are twisted out of the {N(2),C(1),C(7)} least-squares plane by ca. 27° and 47° for C(7) and C(1), respectively. In the other structures the differences between the Nβ-Cipso distances are significantly larger (0.035(3)0.041(4) Å), and at the same time the differences in the extent of twisting of the phenyl rings out of the {N(2),N(1),N(7)} planes become much larger. In 2, 11, and 12 one phenyl ring is much more coplanar (twist angle 4-9°) than the other (twist angle 47-69°). The more coplanar phenyl rings are associated with the shorter Nβ-Cipso distances (range 1.393(3)-1.399(7) Å, cf. 1.433(2)-1.438(7) Å). These trends are evidence for efficient π conjugation between Nβ and the phenyl substituents.83 Previous ab initio calculations have found that the free model hydrazide dianion [NNH2]2- has a pyramidalized β-nitrogen, as does the monoanion and the triplet form of neutral NNH2.76 In contrast, the formally dianionic31 NNPh2 groups in group 4 systems possess planar β-nitrogens. Since terminal TidNNMe2 analogues feature pyramidal β-Ns,30,34 it appears that the planarity of the diphenyl hydrazide β-nitrogens is due to Nβ-Cipso conjugation. It is interesting to note the variation in Ti(1)-N(1), N(1)-N(2), and Ti(1)-N(1)-N(2) parameters on going from 2 to 12 (Table 3). There is an apparent trend in lengthening of both the Ti(1)-N(1) and N(1)-N(2) distances and a decrease in the Ti(1)-N(1)-N(2) angle along this series. At the same time, as mentioned above, the diamide-amine Ti(1)-N(3) and Ti(1)-N(4) distances are decreasing, and the amide nitrogen substituents are becoming more coplanar with the {Ti(1),N(1),N(2),N(3)} (83) Either strongly to just one Ph ring (e.g., 2, 11, 12) or in an intermediate way to both (e.g., 7, 8, 10)

Titanium Hydrazides

Figure 4. σ-Only frontier MOs (d orbitals only, arbitrary energies) for hypothetical bis(dialkylamide)-hydrazide system 17. Labels are for C2V symmetry, and Lax represents a nominal axial ligand.81,82

trigonal plane. In contrast, the orientation of the NβPh2 moiety is preserved in all these compounds, with the {N(1),N(2),C(1),C(7)} least-squares plane being effectively coplanar with {Ti(1),N(1), N(2),N(3)} in all instances. To help account for these observations and to gain a better understanding of their bonding in general, we analyzed the competition between the amide and hydrazide nitrogen π-donor orbitals for the Ti(dπ) acceptor MOs and considered how this would be affected by the orientation of the various amide N and hydrazide Nβ substituents. Molecular Orbital Analysis of Titanium Hydrazides Containing Dianionic N3-Donor Ligands. We have previously reported detailed analyses of the metal(dπ)-N(pπ) bonding in d0 imido complexes of the type M(N2Npy)(NR)(L) (L ) σ donor).81,82 This serves as a starting point for considering the π-bonding situation in 2 and 10-12. Figure 4 illustrates the σ-only frontier MOs (d orbitals only, arbitrary energies) for a hypothetical trigonal-bipyramidal complex M(NNR2)(NR′2)2(Lax)2 (17), where Lax is a σ-only donor. The 2a1 MO is strongly σ* antibonding and would not engage in metal-ligand π interactions. The 1a1 and 1b1 levels are slightly σ* metal-ligand antibonding and higher in energy than the strictly σ-nonbonding 1b2 and 1a2 levels. Although these four orbitals are all available for metal-ligand π-bonding interactions with correctly oriented amide and hydrazide pπ donors, the 1b2 and 1a2 levels are the acceptors with the best energy match. Figure 5a shows schematically the four highest occupied MOs of a planar, dianionic [NNR2]2- ligand based on previous EHMO and DFT calculations.27,31,76,84 The 1a1 MO is suitable for M-NR σ bonding, and the 1b2 MO is NR-Nβ π bonding and would not be expected to interact to any significant extent with a metal center.31,76 The 1b1 MO (which we abbreviate as πh) is essentially NR-Nβ nonbonding and would form one of the M-NR π bonds. The 2b2 MO (abbreviated πv) is the HOMO of [NNR2]2- and lies significantly higher in energy (g1 eV31,76) than the 1b1 MO because of its NR-Nβ π*-antibonding nature. Of these two MOs (πv (2b2) and πh (1b1)), the former is the better π donor due to its better energy match with the metal acceptor orbitals. Donation of electron density from the πv strengthens the NR-Nβ bond due to the formal removal of π*antibonding electron density. The consistent orientation of the (84) Kahlal, S.; Saillard, J.-Y.; Hamon, J.-R.; Manzur, C.; Carrillo, D. New J. Chem. 2001, 25, 231.

Organometallics, Vol. 27, No. 24, 2008 6487

NβPh2 moiety in 2-12 is therefore understood in terms of providing the best interaction between the NNR2 and titanium since this provides the most favorable overlap between the highest energy donor MO of NNPh2 (2b2) and the appropriate lowest energy acceptor orbital at titanium (1b2 in Figure 4). The second π interaction between Ti and NR is formed between the 1b1-donor MO of NNPh2 and the 1b1 acceptor at the metal center. The trigonal planar amide nitrogens of the real compounds 2 and 10-12 suggest that they are able in principle to donate two electrons each from the occupied 2pπ AOs (giving each complex an 18 valence electron count). However, as discussed previously,81,82 the situation is complicated by symmetry restrictions, the details of which depend on the orientation of the lone pairs with respect to the equatorial plane of the complex. In the model complex 17 two extreme positions can be identified as depicted in Figure 5b,c. In Figure 5b the lone pairs are lying in the equatorial plane and their orientation is defined by a dihedral angle NR-Ti-N-R′ of 90°. In Figure 5c the lone pairs are perpendicular to the equatorial plane and the dihedral angle NR-Ti-N-R′ is 0°. In both cases it can be seen that one of the amide SALCs competes with one of the NNR2 MOs of the corresponding symmetry for the same π-acceptor orbital at the metal center. This is either the 1b1 orbital in the case of Figure 5b or the 1b2 orbital when the amide SALCs are oriented as in Figure 5c. Thus in either extreme (and at all points in between) there is always a three-center-four-electron π-bonding conflict, and so the real complexes are best described as 16 valence electron “π-loaded” systems. This type of situation is well established in imido chemistry with very interesting consequences for metal-ligand multiple bonding properties and reactivity patterns.1,81,82,85-90 To test these qualitative ideas, the structure of Ti(NNPh2)(N2NC2,Me)(py) (2) has been optimized using DFT. The agreement between the computed (see Table S3 of the Supporting Information) and experimental geometries is very good. The four highest occupied MOs for 2 are shown in Figure 6. The orientation of the amide donors in 2 lies closer to the situation described by Figure 5b, and the appearance and energies of the HOMOs can be interpreted in terms of the qualitative model described above. The HOMO of 2 is the bonding combination between πv of NNPh2 and 1b2 on Ti and lies 1.9 eV above the corresponding πh-derived orbital (HOMO-3). Both the HOMO and HOMO-3 are heavily ligand-based, and the NNPh2 is best described as a hydrazido(2-) ligand forming a Ti-NR triple bond. The very large destabilization of the HOMO is attributed to the NR-Nβ antibonding contribution. The LUMO of 2 is essentially the LUMO of pyridine with some weak bonding Ti-N interaction with a 3d AO on Ti. The HOMO-3 also contains a minor contribution from the out-of-phase amide nitrogen lone pairs of the N2NC2,Me ligand (formally the b1 SALC in Figure 5b), which is competing with πh for the single suitable acceptor MO on titanium (equivalent (85) Green, J. C.; Green, M. L. H.; James, J. T.; Konidaris, P. C.; Maunder, G. H.; Mountford, P. J. Chem. Soc., Chem. Commun. 1992, 1361. (86) Dunn, S. C.; Mountford, P.; Robson, D. A. J. Chem. Soc., Dalton Trans. 1997, 293. (87) Parkin, G.; van Asselt, A.; Leahy, D. J.; Whinnery, L.; Hua, N. G.; Quan, R. W.; Henling, L. M.; Schaefer, W. P.; Santarsiero, B. D.; Bercaw, J. E. Inorg. Chem. 1992, 31, 82. (88) Schofield, M. H.; Kee, T. P.; Anhaus, J. T.; Schrock, R. R.; Johnson, K. H.; Davis, W. M. Inorg. Chem. 1991, 30, 3595. (89) Benson, M. T.; Bryan, J. C.; Burrell, A. K.; Cundari, T. R. Inorg. Chem. 1995, 34, 2348. (90) Lin, Z.; Hall, M. B. Coord. Chem. ReV. 1993, 123, 149.

6488 Organometallics, Vol. 27, No. 24, 2008

Selby et al.

Figure 5. (a) Four highest occupied MOs of a C2V symmetric dianion [NNR2]2-. (b) Amide lone pair SALCs when the lone pair 2p AOs lie in the equatorial plane. (c) Amide lone pair SALCs when the lone pair 2p AOs are perpendicular to the equatorial plane. Symmetry labels are given for the C2V point group.

Figure 7. Simplified models based on Ti(NR)(NMe2)2(py)2 (R ) NMe2, Me) used to evaluate the effect of changing the orientation or type of various substituents. The suffixes “90” and “0” refer to the magnitude of the torsion angles NR-Ti-Nam-Me (cf. the torsion angles N(1)-Ti(1)-N(3,4)-Si or N(1)-Ti(1)-N(3,4)-C in Table 3). Figure 6. Four highest occupied MOs of Ti(NNPh2)(N2NC2,Me)(py) (2) and their DFT-computed energies.

to 1b1 in Figure 4). The HOMO-2 of 2 represents a Ti-Namide π-bonding interaction from the in-phase amide nitrogen lone pairs of N2NC2,Me (formally the a1 SALC in Figure 5b), which does not compete with either πv or πh. The HOMO-1 is essentially a Ti-N nonbonding, ligand-based lone pair in nature. Its major contribution comes from the out-of-phase amide nitrogen lone pairs of N2NC2,Me, although there is also a contribution from πh. Therefore each amide nitrogen of N2NC2,Me is effectively a two-electron donor as a consequence of the π-loaded nature of 2. It appears that πh competes more effectively than the amide b1 SALC for the single b1-type acceptor orbital at titanium, consistent with the shorter Ti-NR compared to Ti-Nam bond distances (calculated values 1.735 and 2.015 (av) Å, respectively). However, the most remarkable feature of the bonding in 2 is the significant destabilization of the HOMO. This not only lies substantially above the other Ti-NR π bond (HOMO-3) but is even ca. 1 eV above the nonbonding HOMO-1. The high energy of this MO (which is simultaneously Ti-NR bonding and NR-Nβ antibonding) may be responsible for the very novel reactivity found between 2 and alkynes (eq 2). NBO Analysis of Model Hydrazido and Imido Complexes. As mentioned, 2 lies closer to the situation in Figure 5b with the amide lone pairs perpendicular to the equatorial plane. On the other hand, 12 lies closer to that in Figure 5c

(lone pairs lying in the equatorial plane), and on going from one to the other, the Ti-Nam distances progressively shorten while both the TidNR and NR-Nβ lengthen (Table 3). A clear understanding of these effects will be important for understanding the reactivity patterns of such compounds and will aid selection of suitable supporting ligands. Therefore, using a series of simplified models based on Ti(NR)(NMe2)2(py)2 (R ) NMe2, Me; Figure 7), we investigated the differing π-donor effects of the various diamide-amide ligands in 2 and 10-12. We varied the orientation of the β-NMe2 group and compared the bonding between model hydrazide systems (L)TidNNMe2 and the corresponding imides (L)TidNMe. The model compounds are shown in Figure 7 along with the abbreviations used. Table 4 lists key geometrical parameters and relative energies for the various systems. The model systems (Figure 7; see the Computational Details for full details) were constrained to C2V symmetry with planar β-nitrogens in the case of the hydrazides (in the real systems 2 and 10-12 the phenyl substituents stabilize the trigonal-planar Nβ geometry). To understand the structural variations (Table 4) between the different models, we analyzed the electronic structures using the NBO (natural bond orbital) method.91,92 The advantage of this approach is that it aids interpretation of the bonding in a transparent and intuitive way in terms of the closest (91) Reed, A. E.; Curtiss, L. A.; Weinhold, F. Chem. ReV. 1988, 88, 899.

Titanium Hydrazides

Organometallics, Vol. 27, No. 24, 2008 6489

Table 4. Key Geometrical Parameters and Relative Energies (Erel) for the Various Model Systems Ti(NR)(NMe2)2(py)2 (R ) NMe2, Me) Studieda parameter

Hyd_90

Hyd_0

Hyd_0_rot

TidNR (Å) NR-Nβ (Å) Ti-Nam (Å) Ti-py (Å) NR-Ti-Nam (deg) Nam-Ti-Nam (deg) Erel (kJ mol-1)

1.734 1.331 2.048 2.288 114.4

1.729 1.341 1.975 2.306 114.2

1.719 1.355 1.980 2.317 110.6

Imide_90 1.705

Imide_0 1.708

2.050 2.269 114.2

1.970 2.307 113.7

131.1

131.5

138.8

131.8

132.5

72.2

0.0

33.6

93.6

0.0

a The energies for Hyd_90 and Hyd_0_rot are expressed relative to that of Hyd_0. The Erel of Imide_90 is expressed relative to that of Imide_0. Nam refers to the NMe2 amide nitrogens. See Figure 7 for the geometries.

idealized Lewis structure. For all of the models Ti(NR)(NMe2)2(py)2 the NBO analysis found NLMOs (natural localized molecular orbitals) corresponding to Ti-NR and Ti-Nam σ bonds, two Ti-NR π bonds (one for each of πv and πh), and a Ti-Nam π bond to each NMe2. For the hydrazide models a “lone pair” essentially localized (see below) on Nβ was also found. These NLMOs clearly correspond to the expected Lewis structures 18 and 19 for the hydrazido(2-) and imido models, respectively. The main % atomic contributions to these NLMOs are given in Table 5.

Orientation of Amide Lone Pairs. Going from model Hyd_90 to Hyd_0 rotates the NMe2 lone pairs from lying in the equatorial plane to being perpendicular to it. This models the general trend on going from 2 to 12 in the real systems. For the model systems this results in a very slight shortening of TidNR, a significant shortening of Ti-Nam, and a lengthening of NR-Nβ. There is an associated stabilization of 72.2 kJ mol-1. The shorter Ti-Nam may be attributed to the better match between the amide lone pairs and the 1b2 and 1a2 acceptor MOs of Ti (cf. Figure 4) than the 1a1 and 1b1 alternatives. Consistent with this, the NBO analysis (Table 5) shows an increased covalency in both the σ and π Ti-Nam NLMOs.93 The competition between the amide lone pairs and the hydrazide πv for the Ti 1b2 acceptor MO results in a diminished πv(Ti-NR) interaction. Conversely, the corresponding πh(Ti-NR) bonding is improved since the Ti 1b1 no longer overlaps with the amide lone pairs. Importantly, the higher %Ti in σ(Ti-NR) shows that the NNMe2 ligand is a better σ donor in Hyd_0 because there is no longer destabilization of the Ti 1a1 orbital (Ti-NR σ* antibonding) through π donation from the amide a1 lone pair. Thus there is a combination of σ and π effects that contribute (92) Weinhold,F.; Landis, C. R. Valency and Bonding: A Natural Bond Orbital Donor-Acceptor PerspectiVe; Cambridge University Press: Cambridge, 2005. (93) Because of the higher electronegativity of nitrogen, all of the occupied NLMOs have a higher %N contribution than %Ti. Therefore an increased %Ti and decreased %N contribution for a given NLMO on going from one model to another is evidence of improved covalency in the orbital under consideration. From the perspective of fragment MO interactions this relates to a decreased energy separation and/or improved overlap between the participating orbitals.

to the stabilization of isomer Hyd_0 compared with Hyd_90, and the variation in the TidNR distance is a composite of increased σ(NR) and πh donation and diminished πv donation. We have also calculated the NPA (natural population analysis) charges for the fragments {NR}, {(NMe2)2} and {Ti(py)2} for the five model systems Ti(NR)(NMe2)2(py)2 (see Table S4 of the Supporting Information). On going from Hyd_90 to Hyd_0, the {Ti(py)2} fragment charge decreases from +1.753e to +1.585e while the {(NMe2)2} charge changes from -1.173e to -1.000e. The {NNMe2} charge on the other hand decreases by only 0.005e. Similar variations were found for the imido systems and again illustrates the effect of rotating the amide lone pairs on NMe2fTi electron donation. Rotation of the NβMe2 Group. All of the experimental systems 2-12 have the hydrazide NPh2 substituents lying in the equatorial plane. As discussed above, this should allow best overlap of the hydrazide πv orbital with the lower energy acceptor orbital on Ti. To quantify this analysis, we examined the model Hyd_0_rot (Figure 7), which has the NβMe2 group rotated by 90°. This was accompanied by a destabilization (∆Erel) of 33.6 kJ mol-1 (Table 4) relative to Hyd_0, consistent with experimental observations. The NLMO analysis for Hyd_0_rot is similar to that for Hyd_0 with regard to the Ti-Nam interactions and Ti-NR σ interactions. However, there is a significant effect on the Ti-NR π interactions as expected. Whereas the πv(Ti-NR) bond is significantly more covalent than πh(Ti-NR) in Hyd_0 (and also Hyd_90) for the reasons discussed above, in Hyd_0_rot the Ti-NR π bonding is much more cylindrical. This is because the higher energy NNMe2 π donor (πv) now has to interact with a higher energy Ti dπ acceptor (1b1, which is also slightly Ti-Nam σ* antibonding), while the πh donor accesses the lower energy Ti 1b2 orbital. Overall there is a net gain in negative charge of the {NNMe2} and {(NMe2)2} fragments (Table S4 in the SI) and increase in positive charge for {Ti(py)2}, showing that this orientation of the NβMe2 group results in a reduced net donation from both the diamide and hydrazide ligands, despite the slight shortening of the Ti-NR bond length. Nr-Nβ Interaction. The weakening of the NR-Nβ bond is an important aspect of metal hydrazide chemistry from the point of view of its eventual cleavage. Several reports from group 4 have found facile cleavage of this bond, especially in π-loaded systems.29,34,35 Experimentally (2-12) and computationally it is found that rotation of the amide lone pairs results in a slightly longer NR-Nβ distance (∆ ≈ 0.01 Å). Furthermore, the model Hyd_0_rot shows a further ca. 0.015 Å lengthening of NR-Nβ and shortening of Ti-NR. The NBO and NLMO analyses offer an explanation of these trends in Ti-NR and NR-Nβ distances. As mentioned, the NBO analyses of Hyd_90, Hyd_0, and Hyd_0_rot are consistent with the Lewis structure 18, in which there is a formal “lone pair” mainly localized on Nβ. This is confirmed by the NLMO analysis for this orbital (Table 5; %N ca. 87-89%). However, this also finds that the electron pair is partially delocalized into the antibonding π*v(Ti-NR) orbital, accounting for the nonzero %Ti and %NR contributions to the “lone pair” NLMO (see Figure S2 of the Supporting Information). Clearly, this Nβfπ*v(Ti-NR) interaction therefore simultaneously weakens the Ti-NR bond and strengthens the NR-Nβ bond. The extent of the interaction will vary according to the properties (energy and %NR character) of the π*v(Ti-NR) orbital. On going from Hyd_90 to Hyd_0, the π*v(Ti-NR) orbital is destabilized due to competition with the amide lone pairs for the Ti 1b2 acceptor orbital, and in Hyd_0_rot the NNMe2 πv donor has to use the higher energy 1b1 orbital.

6490 Organometallics, Vol. 27, No. 24, 2008

Selby et al.

Table 5. NLMO Analysis for the Various Model Systems Ti(NR)(NMe2)2(py)2 (R ) NMe2, Me) Studieda electron pair σ(Ti-NR) πv(Ti-NR) πh(Ti-NR) “lone pair” (Nβ) σ(Ti-Nam) π(Ti-Nam)

Hyd_90

Hyd_0

Hyd_0_rot

Imide_90

Imide_0

11.9% Ti 86.2% NR 36.3% Ti 62.1% NR 16.7% Ti 80.7% NR 5.2% Ti 3.3% NR 86.7% Nβ 9.2% Ti 87.9% Nam 8.9% Ti 85.2% Nam

15.7% Ti 83.6% NR 32.7% Ti 65.6% NR 18.4% Ti 78.8% NR 4.5% Ti 2.5% NR 88.3% Nβ 11.9% Ti 85.6% Nam 9.2% Ti 83.6% Nam

15.8% Ti 83.0% NR 28.5% Ti 69.5% NR 22.5% Ti 75.3% NR 3.6% Ti 1.8% NR 89.0% Nβ 11.3% Ti 86.1% Nam 9.8% Ti 83.1% Nam

16.1% Ti 83.2% NR 27.6% Ti 69.9% NR 21.2% Ti 76.3% NR

18.4% Ti 81.2% NR 25.4% Ti 72.0% NR 22.0% Ti 75.2% NR

9.0% Ti 87.3% Nam 8.1% Ti 85.7% Nam

12.0% Ti 85.4% Nam 8.8% Ti 83.4% Nam

a The % contributions are for the stated atoms in the respective NLMOs. See Figure 7 for the geometries. NPA (natural population analysis) charges for the {Ti(py)2}, {NR}, and {(NMe2)2} fragments are given in Table S4 of the Supporting Information.

Table 6. X-ray Data Collection and Processing Parameters for Ti(NNPh2)Cl2(py)3 (7), Ti(NNPh2)Cl2(py′)3 (8), Ti(NNPh2)(N2NC2,Me)(py) (2), Ti(NNPh2)(N2NC2,SiMe3)(py) (10), Ti(NNPh2)(N2NC3,Me)(py) (11), and Ti(NNPh2)(N2Npy)(py) (12) empirical formula fw temp/K wavelength/Å space group a/Å b/Å c/Å R/deg β/deg γ/deg V/Å3 Z d(calcd)/Mg · m-3 abs coeff/mm-1 R indices: R1 ) [I > 3σ(I)]:a Rw ) a

7

8

2

10

11

12

C27H25Cl2N5Ti 538.33 150 0.71073 C2/c 14.0110(4) 19.8007(6) 9.4425(3) 90 93.9572(12) 90 2613.36(14) 4 1.368 0.557 0.0337 0.0367

C39H49Cl2Ti 706.66 150 0.71073 P21/c 22.2899(2) 17.6264(2) 19.8028(2) 90 91.9807(6) 90 7775.69(14) 8 1.207 0.390 0.0397 0.0474

C28H44N6Si2Ti 568.77 150 0.71073 P1 10.1794(5) 11.1800(7) 13.8952(8) 89.726(3) 77.178(2) 85.912(3) 1537.9(2) 2 1.228 0.383 0.0728 0.0800

C30H50N6Si3Ti 626.93 150 0.71073 P1j 10.3113(2) 11.9118(3) 14.0949(3) 92.1330(10) 90.7564(10) 90.6550(10) 1729.75(7) 2 1.204 0.380 0.0350 0.0369

C30H48N6Si2Ti 596.82 150 0.71073 P 21/c 11.5182(2) 14.5639(3) 20.0751(4) 90 100.6813(8) 90 3309.25(11) 4 1.198 0.359 0.0429 0.0387

C32H44N6Si2Ti 616.81 150 0.71073 P1j 10.1894(2) 10.3302(2) 17.9754(4) 95.5849(10) 103.2464(11) 111.5826(10) 1677.77(6) 2 1.221 0.357 0.0524 0.0448

R1 ) ∑||Fo| - |Fc||/∑|Fo|; Rw ) {∑w(|Fo| - |Fc|)2/∑w|Fo|2}.

Therefore the Nβfπ*v(Ti-NR) interaction is diminished along this distortion coordinate, consistent with the computed bond length trends. Indeed, the second-order perturbation energy associated with the Nβfπ*v(Ti-NR) interaction decreases from Hyd_90 to Hyd_0 and finally to Hyd_0_rot (141.9, 124.1, and 103.4 kJ mol-1, respectively) due to both a larger energy gap (0.26, 0.27, and 0.28 au, respectively) and a smaller Fock matrix element F(i,j) (0.086, 0.080, and 0.074 au, respectively). This further explains the trends in the composition of the lone pair on Nβ in Table 5. Comparison with Imido Systems. The NLMO analysis for the imido systems Imide_90 and Imide_0 reveals analogous trends, and in fact the latter is 93.6 kJ mol-1 more stable. The Ti-Nam distances shorten significantly (∆(Ti-Nam) ) -0.08 Å), whereas (as for the hydrazido compounds) no significant change is registered for Ti-NR. These small changes in Ti-NR for the hydrazido and imido systems show there is no significant net electronic driving force for Ti-NR bond length change, despite varying the supporting diamide ligand set. Examination of the πv(Ti-NR) and πh(Ti-NR) NLMOs shows a much more cylindrical π-bonding manifold than is found in the corresponding hydrazide systems. As expected, the πh(Ti-NR) NLMO is more based on NR owing to the higher energy of the Ti 1b1 acceptor MO. The Ti-NR bond distances in Imide_90 and Imide_0 are ca. 0.020-0.030 Å shorter than the corresponding values for Hyd_90 and Hyd_0. This is consistent with the differences found in the real systems Ti(“N2N”)(NR)(py) (“N2N” ) N2NC2,SiMe3 or N2Npy; R ) tBu or NNPh2; ∆(Ti-NR) )

0.032(4)-0.035(3) Å).69,81 One Ti-NR bond-lengthening effect is due to the Nβfπ*v(Ti-NR) interaction noted above. A second contribution is the reduced σ-donor ability of NNR2 compared to NR because of the increased electron-withdrawing inductive effect of NβR2 compared to alkyl. This is evident in the %Ti,N contributions for the σ(Ti-NR) NLMOs in the two types of model system. The NLMOs for the hydrazido models are more localized on NR in each case compared to the imido analogues. A similar conclusion with regard to σ-donor abilities was reached previously (based on fragment MO Mulliken population analyses) in comparing the Ti-NR bonding in the model complexes Ti(NMe){HC(pz)3}Cl2 and Ti(NMe){HC(pz)3}Cl2.31 Comparison of Model and Experimental Systems. The shortening of the av Ti-Nam distances from 2.028 to 1.974 Å from 2 to 12 (Table 3) is reproduced well by the models Hyd_90 (Ti-Nam ) 2.048 Å) and Hyd_0 (Ti-Nam ) 1.975 Å). The increase in the NR-Nβ distance is also well reproduced (0.01 Å in both experiment and theory). However, on going from 2 to 12 in the experimental systems, the Ti-NR distance lengthens from 1.733(5) to 1.759(2) Å, whereas in the model systems there is a very small decrease from 1.734 to 1.729 Å. This apparent discrepancy can be attributed to significant steric interactions between the hydrazide phenyl substituents and the equatorial NSiMe3 groups, as the latter move up into the equatorial plane and become more oriented toward the NNPh2. This effect is probably the origin of the relatively large jump of ca. 0.02 Å in Ti-NR bond length between Ti(NNPh2)(N2NC2,Me)(py) (2) and Ti(NNPh2)(N2NC2,SiMe3)(py) (10) as the diamide-amine

Titanium Hydrazides

Organometallics, Vol. 27, No. 24, 2008 6491

N-Me group is replaced by SiMe3. Thus a balance of real system steric factors and underlying electronic factors leads to the observed structural trends and (most likely) controls the reactivity patterns.

Conclusions Far from being simple “imido analogues”, terminal group 4 hydrazides can exhibit novel and potentially catalytic reactivity of the MdN-NR2 functional group.29,34,35,94,95 In this contribution, based on new readily available synthons, we have reported a comprehensive study of the synthesis and bonding of a new family of titanium diphenyl hydrazides using the important diamide-amine class of supporting ligands, among others. Through this series, the ligand coordination number, steric requirements, chelate ring size, and amide π-donor ability may be controlled. Detailed structural and computational studies of a series of closely related model complexes have defined the parameters, both steric and electronic, which influence the energies of these systems and the Ti-Nam, NR-Nβ, and Ti-NR distances. Depending on their relative orientations, competition between the amide and hydrazide nitrogen 2pπ lone pairs modifies the πv(Ti-NR) NLMOs, which in turn leads to poorer NR-Nβ π bonding and a less stabilized β-nitrogen. The weakening of the NR-Nβ bond through NamfTi π donation to the relevant dπ AO echoes the dramatic weakening that occurs in the Schrock and Chatt catalytic cycles on addition of an external electron, just prior to proton-induced cleavage. Ligands that deploy steric bulk in the vicinity of the TidNNPh2 functional group can also lead to lengthened Ti-NR bonds. Our synthetic and computational results also shed light on the high reactivity so far found for the terminal zirconium hydrazides Cp2Zr(NNPh2)(DMAP)29 and Zr(N2Npy*)(NNPh2)(py), which have reactive NR-Nβ bonds that undergo unique insertion reactions.35 In these systems we again expect the inherent “π-loaded” nature96 to result in destabilized πv(Zr-NR) NLMOs just as for the diamide-amine-supported titanium systems described above. Furthermore, the less stabilized 4dπ AO of Zr is also expected to lead to poorer Nβfπ*v(Zr-NR) interactions. Indeed, the NR-Nβ bond distance of 1.398(3) Å in Zr(N2Npy*)(NNPh2)(py), which also features bulky SiMe2tBu substituents, is the longest known for a terminal diphenyl hydrazide. Finally we recall that terminal metal hydrazides are commonly first-formed products in the functionalization of molecular dinitrogen.7-12 Incorporating some of these above-mentioned design principles into transition metal compounds suitable for N2 activation may lead ultimately to new dinitrogen functionalization processes via activated terminal hydrazides.

Experimental Section General Methods and Instrumentation. All manipulations were carried out under an inert atmosphere of argon or dinitrogen using standard Schlenk-line or drybox procedures. Solvents were predried over activated 4 Å molecular sieves and refluxed over sodium (toluene), sodium/potassium (pentane, diethyl ether), potassium (tetrahydrofuran), or calcium hydride (dichloromethane) under a dinitrogen atmosphere and collected by distillation. Alternatively, solvents were degassed by sparging with dinitrogen and dried by passing through a column of activated alumina. Deuterated solvents (94) Herrmann, H.; Wadepohl, H.; Gade, L. H. Dalton Trans. 2008, 2111. (95) Mindiola, D. J. Angew. Chem., Int. Ed. 2008, 47, 1557. (96) Green, J. C. Chem. Soc. ReV. 1998, 27, 263.

were dried over potassium (C6D6) or P2O5 (CD2Cl2), distilled under reduced pressure, and stored under dinitrogen in Young’s Teflon valve ampules. Solution NMR samples were prepared under a dinitrogen atmosphere in a drybox, in 5 mm Wilmad NMR tubes possessing Young’s Teflon valves. 1H and 13C NMR spectra were recorded on a Varian Mercury 300 spectrometer at ambient temperature. All spectra were internally referenced to either residual protio-solvent (1H) or solvent (13C) resonances and are reported relative to tetramethylsilane (δ ) 0 ppm). Chemical shifts are quoted in δ (ppm) and coupling constants in Hz. Where necessary, 1H and 13C assignments were assisted by the use of two-dimensional 1H-1H and 1H-13C correlation experiments. IR spectra were recorded on a Perkin-Elmer 1710 FTIR spectrometer. Samples were prepared in a drybox as Nujol mulls between NaCl plates. IR data are quoted as wavenumbers (cm-1) within the range 4000-400 cm-1. Mass spectra were recorded by the departmental service, and elemental analyses were carried out by the Elemental Analysis Service at the London Metropolitan University. Starting Materials. The compounds Ti(NtBu)Cl2(py)3,66 [Ti(NNMe2)Cl2(py)2]2,31 Li2N2NC2,SiMe3,98 Li2N2NC3,Me,97 Li2N2NC3,Me,97 Li2N2Npy,99 Li2N2NN′,78 Li2Me4taa,100 and Na2O2NN′ · 0.16(THF)101 were prepared according to the literature methods. Ti(NtBu)Cl2(py′)3 was prepared by analogy with the literature methods for Ti(NtBu)Cl2(py)3 and Ti(NtBu)Cl2(py′)2.66,74 Diphenylhydrazine was obtained from Sigma-Aldrich as the hydrogen chloride salt, from which the free hydrazine was obtained by basification, drying, and removal of residual solvent, followed by distillation under inert atmospheric conditions. Pyridine was dried over freshly ground CaH2 and distilled before use. Other reagents were obtained commercially and used as received. Ti(NNPh2)Cl2(py)3 (7). To a stirred solution of Ti(NtBu)Cl2(py)3 (5.00 g, 11.7 mmol) in benzene (50 mL) was added H2NNPh2 (2.15 g, 11.7 mmol) in benzene (10 mL) dropwise over 15 min. The orange solution was stirred for 14 h, resulting in a dark brown solution and a yellow precipitate. The solution was filtered, and the precipitate washed with pentane (3 × 20 mL) and filtered. The yellow solid product was dried in Vacuo. Yield: 5.60 g (89%). Diffraction-quality crystals were grown from a saturated benzene solution. 1 H NMR (CD2Cl2, 299.9 MHz, 293 K): δ 9.01 (4 H, d, 3J ) 5 Hz, o-H of cis-NC5H4), 8.59 (2 H, br m, coupling not resolved, o-H of trans-NC5H4), 7.86 (2 H, t, 3J ) 7.7 Hz, p-H of cis-NC5H4), 7.68 (1 H, br m, coupling not resolved p-H of trans-NC5H4), 7.41 (4 H, d, 3J ) 7.3 Hz, o-C6H5), 7.29-7.13 (10 H, overlapping peaks, m-C6H5, m-H of cis- and trans-NC5H4), 6.99 (2 H, t, 3J ) 7.3 Hz, p-C6H5). IR (NaCl plates, Nujol mull, cm-1): ν 1653 (w), 1604 (s), 1594 (s), 1586 (s), 1559 (w), 1543 (m), 1487 (m), 1447 (s), 1337 (w), 1313 (w), 1269 (s), 1219 (m), 1168 (w), 1154 (w), 1108 (w), 1070 (m), 1043 (m), 1028 (w), 1013 (m), 842 (w), 753 (m), 693 (s), 636 (w). Anal. Found (calcd for C27H25Cl2N5Ti): C, 60.0 (60.2); H, 4.8 (4.7); N, 12.9 (13.0). A 13C NMR spectrum could not be obtained due to the low solubility of 7 in suitable solvents. Alternative Synthesis of [Ti(NNPh2)Cl2(py)2]2 (4). Ti(NNPh2)Cl2(py)3 (7, 0.50 g, 0.93 mmol) was held under a dynamic vacuum (4 × 10-2 mbar) for 14 h, then washed with pentane (2 × 25 mL) to quantitatively yield [Ti(NNPh2)Cl2(py)2]2 (97) Ward, B. D.; Dubberley, S. R.; Maisse-Francois, A.; Gade, L. H.; Mountford, P. Dalton Trans. 2002, 4649. (98) Cloke, F. G. N.; Hitchcock, P. B.; Love, J. B. J. Chem. Soc., Dalton Trans. 1995, 25. (99) Galka, C. H.; Tro¨sch, D. J. M.; Schubart, M.; Gade, L. H.; Radojevic, S.; Scowen, I. J.; McPartlin, M. Eur. J. Inorg. Chem. 2000, 2577. (100) Black, D. G.; Swenson, D. C.; Jordan, R. F.; Rogers, R. D. Organometallics 1995, 14, 3539. (101) Toupance, T.; Dubberley, S. R.; Rees, N. H.; Tyrrell, B. R.; Mountford, P. Organometallics 2002, 21, 1367.

6492 Organometallics, Vol. 27, No. 24, 2008 (4) as a yellow-green solid. The NMR data for 4 were consistent with those previously reported.31 Ti(NNPh2)Cl2(py′)3 (8). To a stirred solution of Ti(NtBu)Cl2(py′)3 (2.00 g, 3.37 mmol) in benzene (50 mL) was added H2NNPh2 (0.62 g, 3.37 mmol) in benzene (10 mL) dropwise over 15 min. The orange solution was stirred for 14 h, resulting in a dark brown solution. Volatiles were removed under reduced pressure, and the dark brown solid was washed with pentane (3 × 20 mL) and filtered. The yellow solid product was dried in Vacuo. Yield: 1.75 g (74%). Diffraction-quality crystals were grown from a saturated pentane solution. H NMR (C6D6, 299.9 MHz, 293 K): δ 9.34 (4 H, d, 3J ) 6.5 Hz, o-H of cis-NC5H4), 9.03 (2 H, d, 3J ) 4.7 Hz, o-H of transNC5H4), 8.19 (4 H, d, 3J ) 7.7 Hz, o-C6H5), 7.29 (4 H, dd, 3J ) 7.7 and 7.0 Hz, m-C6H5), 6.88 (2 H, t, 3J ) 7.0 Hz, p-C6H5), 6.65 (2 H, d, 3J ) 4.7 Hz, m-H of trans-NC5H4), 6.60 (4 H, d, 3J ) 6.5 Hz, m-H of cis-NC5H4), 0.90 (9 H, s, trans-CMe3), 0.81 (18 H, s, cis-CMe3). 13C{1H} NMR (C6D6, 75.4 MHz, 293 K): δ 162.2 (p-C of cis-NC5H4), 160.1 (p-C of trans-NC5H4), 152.9 (o-C of cisNC5H4), 152.2 (o-C of trans-NC5H4), 145.8 (i-C6H5), 129.4 (mC6H5),123.3 (p-C6H5), 121.2 (m-C of cis-NC5H4), 120.6 (m-C of trans-NC5H4), 120.0 (o-C6H5), 34.9 (cis-CMe3), 34.7 (trans-CMe3), 30.5 (trans-CMe3), 30.3 (cis-CMe3). IR (NaCl plates, Nujol mull, cm-1): ν 1613 (s), 1585 (s), 1543.7 (s), 1461 (s), 1418 (m), 1368 (m), 1314 (s), 1274 (s), 1230 (s), 1201 (m), 1184 (w), 1172 (w), 1070 (s), 1022 (m), 995 (w), 843 (s), 803 (s), 750 (s), 725 (s) 695 (m), 637 (w), 572 (w). Anal. Found (calcd for C39H49Cl2N5Ti): C, 66.4 (66.3); H, 7.1 (7.0); N, 9.8 (9.9). Ti(NNPh2)(N2NC2,Me)(py) (2). To a stirred mixture of Ti(NNPh2)Cl2(py)3 (2.00 g, 3.72 mmol) and Li2N2NC2,Me (1.01 g, 3.72 mmol) was added toluene (60 mL), all at -78 °C. The green suspension was allowed to warm to room temperature and stirred for a further hour, resulting in a dark brown solution. Volatiles were removed under reduced pressure, and the dark oily solid was extracted into Et2O (3 × 30 mL) and filtered twice. Volatiles were removed under reduced pressure to yield a brown solid, which was washed with pentane (3 × 5 mL) cooled to -78 °C. The green solid product was dried in Vacuo. Yield: 1.37 g (65%). Diffractionquality crystals were grown from a saturated pentane solution at -30 °C. 1

1 H NMR (C6D6, 299.9 MHz, 293 K): δ 8.68 (2 H, d, 3J ) 6.4 Hz, o-NC5H5), 7.87 (4 H, d, 3J ) 7.6 Hz, o-C6H5), 7.31 (4 H, app. t, app. 3J ) 7.1 and 7.6 Hz, m-C6H5), 6.91 (2 H, t, 3J ) 7.1 Hz, p-C6H5), 6.68 (1 H, t, 3J ) 7.6 Hz, p-NC5H5), 6.38 (2 H, app. t, app. 3J ) 6.4 and 7.6 Hz, m-NC5H5), 3.56, 3.39 (2 × 2 H, 2 × m, CH2NSiMe3), 2.74, 2.56 (2 × 2 H, 2 × m, CH2NMe), 2.62 (3 H, s, NMe), 0.03 (18 H, s, SiMe3). 13C{1H} NMR (C6D6, 75.4 MHz, 293 K): δ 153.9 (o-NC5H5), 147.9 (i-C6H5), 138.7 (p-NC5H5), 129.3 (m-C6H5), 124.0 (m-NC5H5), 122.2 (p-C6H5), 119.1 (o-C6H5), 63.6 (CH2NMe), 50.7 (NMe), 48.2 (CH2NSiMe3), 1.9 (SiMe3). IR (NaCl plates, Nujol mull, cm-1): ν 1600 (m), 1586 (s), 1486 (s), 1445 (s), 1308 (w), 1298 (w), 1240 (s), 1211 (w), 1168 (w), 1078 (m), 1064 (m), 1041 (m), 1012 (w), 989 (w), 919 (s), 841 (s), 799 (m), 745 (s), 697 (s), 636 (w), 621 (w). EI-MS: m/z 489 (16%) [M py]+, 168 (73%) [NPh2]+. Anal. Found (calcd for C28H44N6Si2Ti): C, 59.0 (59.1); H, 7.7 (7.8); N, 14.7 (14.8). Ti(NNPh2)(N2NC2,SiMe3)(py) (10). To a stirred mixture of [Ti(NNPh2)Cl2(py)2]2 (1.00 g, 2.18 mmol) and Li2N2NC2,SiMe3 (0.72 g, 2.18 mmol) was added toluene (60 mL), all at -78 °C. The green suspension was allowed to warm to room temperature and stirred for a further 90 min, resulting in a dark brown solution. Volatiles were removed under reduced pressure, and the dark oily solid was extracted into Et2O (2 × 30 mL) and filtered. Volatiles were removed under reduced pressure to yield a brown solid, which was washed with pentane (3 × 5 mL) cooled to -78 °C. The green solid product was dried in Vacuo. Yield:

Selby et al. 0.36 g (26%). Diffraction-quality crystals were grown from a saturated pentane solution at -30 °C. 1 H NMR (C6D6, 299.9 MHz, 293 K): δ 8.66 (2 H, d, 3J ) 6.4 Hz, o-NC5H5), 7.78 (4 H, d, 3J ) 8.3 Hz, o-C6H5), 7.28 (4 H, app. t, app. 3J ) 7.1 and 8.3 Hz, m-C6H5), 6.89 (2 H, t, 3J ) 7.1 Hz, p-C6H5), 6.72 (1 H, t, 3J ) 7.6 Hz, p-NC5H5), 6.38 (2 H, app. t, app. 3J ) 6.4 and 7.6 Hz, m-NC5H5), 3.76 (2 H, m, CH2NSiMe3 inner protons cis to py), 3.21 (2 H, m, CH2NSiMe3 outer protons cis to py), 3.06 (2 H, m, CH2NSiMe3 inner protons trans to py), 2.46 (2 H, m, CH2NSiMe3 outer protons trans to py), 0.22 (9 H, s, SiMe3 trans to py), 0.03 (18 H, s, SiMe3 cis to py). 13C{1H} NMR (C6D6, 75.4 MHz, 293 K): δ 152.5 (o-NC5H5), 147.7 (i-C6H5), 138.4 (p-NC5H5), 129.2 (m-C6H5), 123.9 (m-NC5H5), 122.4 (p-C6H5), 119.9 (o-C6H5), 59.5 (CH2NSiMe3 trans to py), 48.7 (CH2SiMe3 cis to py), 1.8 (SiMe3 cis to py), 0.6 (SiMe3 trans to py). IR (NaCl plates, Nujol mull, cm-1): ν 1600 (m), 1583 (m), 1489 (m), 1444 (s), 1345 (w), 1306 (w), 1295 (w), 1276 (m), 1256 (m), 1245 (m), 1211 (w), 1167 (w), 1067 (m), 1041 (w), 1024 (m), 1012 (w), 993 (w), 939 (s), 908 (w), 872 (m), 840 (s), 797 (w), 779 (m), 754 (w), 745 (m), 696 (m), 678 (w), 635 (w), 624 (w). EI-MS: m/z 168 (98%) [NPh2]+, 77 (77%) [C6H6]+. Anal. Found (calcd for C30H50N6Si3Ti): C, 57.4 (57.5); H, 7.9 (8.0); N, 13.4 (13.4). Ti(NNPh2)(N2NC3,Me)(py) (11). To a stirred mixture of [Ti(NNPh2)Cl2(py)2]2 (1.00 g, 2.18 mmol) and Li2N2NC3,Me (0.657 g, 2.18 mmol) was added toluene (60 mL), all at -78 °C. The green suspension was allowed to warm to room temperature and stirred for a further hour, resulting in a dark brown solution. Volatiles were removed under reduced pressure, and the dark oily solid was extracted into Et2O (2 × 30 mL) and filtered. Volatiles were removed under reduced pressure to yield a brown solid, which was washed with hexane (3 × 5 mL) cooled to -78 °C. The brown solid product was dried in Vacuo. Yield: 0.53 g (41%). Diffraction-quality crystals were grown from a saturated hexane solution at -30 °C.

H NMR (C6D6, 299.9 MHz, 293 K): δ 8.66 (2 H, d, 3J ) 6.4 Hz, o-NC5H5), 7.67 (4 H, d, 3J ) 7.6 Hz, o-C6H5), 7.25 (4 H, app. t, app. 3J )7.6 and 7.8 Hz, m-C6H5), 6.90 (3 H, m, p-C6H5 and p-NC5H5), 6.61 (2 H, app. t, app. 3J ) 6.4 and 7.6 Hz, m-NC5H5), 3.93 (2 H, m, CH2NSiMe3 inner protons), 3.28 (2 H, m, CH2NSiMe3 outer protons), 2.55 (3 H, s, NMe), 2.23 (2 H, m, CH2NMe inner protons), 1.99, (2 H, m, CH2NMe outer protons), 1.33 (4 H, m, 2 × CH2), 0.23 (18 H, s, SiMe3). 13C{1H} NMR (C6D6, 75.4 MHz, 293 K): δ 151.8 (o-NC5H5), 147.6 (i-C6H5), 136.6 (p-NC5H5), 129.3 (m-C6H5), 123.7 (m-NC5H5), 122.3 (p-C6H5), 119.3 (o-C6H5), 60.6 (CH2NMe), 50.7 (NMe), 46.5 (CH2NSiMe3), 30.8 (CH2), 2.5 (SiMe3). IR (NaCl plates, Nujol mull, cm-1): ν 1596 (m), 1585 (m), 1488 (s), 1457 (s), 1442 (m), 1297 (m), 1255 (m), 1245 (s), 1209 (w), 1158 (w), 1146 (w), 1135 (m), 1078 (m), 1061 (m), 1036 (m), 1013 (m), 991 (w), 961 (w), 947 (m), 920 (w), 888 (m), 879 (m), 854 (s), 828 (s), 797 (m), 776 (m), 756 (m), 745 (s), 695 (s), 662 (w), 625 (m), 614 (m). EI-MS: m/z 517 (11%) [M - py]+, 168 (100%) [NPh2]+. Anal. Found (calcd for C30H48N6Si2Ti): C, 60.4 (60.4); H, 8.2 (8.1); N, 13.9 (14.1). Ti(NNPh2)(N2Npy)(py) (12). To a stirred slurry of Ti(NNPh2)Cl2(py)3 (1.67 g, 3.10 mmol) in toluene (50 mL), cooled to -78 °C, was added Li2N2Npy (1.00 g, 3.10 mmol) in toluene (10 mL) dropwise over 15 min. The green-yellow slurry was allowed to warm to room temperature and stirred for 14 h, resulting in a dark brown solution. Volatiles were removed under reduced pressure, and the dark brown solid was extracted into Et2O (3 × 30 mL) and filtered. Volatiles were removed under reduced pressure to yield a brown solid, which was stirred with pentane (2 × 50 mL) and filtered. The dark green solid product was dried in Vacuo. Yield: 1.40 g (73%). Diffraction-quality crystals were grown from slow evaporation of a diethyl ether solution. 1

1

H NMR (C6D6, 299.9 MHz, 293 K): δ 9.44 (1H, d, py-H6, 3J ) 4.1 Hz), 9.00 (2H, d, o-NC5H4, 3J ) 4.1 Hz), 7.82 (4 H, d, 3J ) 7.6 Hz, o-C6H5), 7.29 (4 H, app. t, app. 3J )7.6 and 8.3 Hz,

Titanium Hydrazides m-C6H5), 7.03 (1 H, ddd, py-H4, 3J ) 7.6 Hz, 4J ) 1.8 Hz) 6.93-6.85 (4 H, overlapping m, p-C6H5, p-NC5H4, and py-H3), 6.61 (2 H, app t, m-NC5H4, app. 3J ) 6.8 Hz), 6.41 (1 H, dd, pyH5, 3J ) 7.6 Hz, 4J ) 1.2 Hz), 3.79 (2 H, d, CHHNSi3, 2J ) 12.3 Hz), 3.37 (2 H, d, CHHNSi3, 2J ) 12.3 Hz), 1.20 (3 H, s, Me), -0.03 (18 H, s, NSiMe3). 13C{1H} NMR (C6D6, 75.4 MHz, 293 K): δ 161.7(py-C2), 152.4 (o-NC5H4), 151.9 (py-C6), 148.1 (i-C6H5), 138.3 (py-C4), 137.6 (p-NC6H5) 129.6 (m-C6H5), 123.8 (m-NC5H4), 122.2 (p-C6H5), 121.3 (py-C5), 119.7 (py-C3), 119.5 (o-C6H5), 64.3 (CH2NSi3), 47.3(MeC), 24.6 (MeC), 1.3 (NSiMe3). IR (NaCl plates, Nujol mull, cm-1): ν 1596 (s), 1585 (s), 1559 (w), 1506 (s), 1486 (s), 1467 (m), 1313 (s), 1299 (m), 1282 (s), 1240 (s), 1212 (m), 1159 (w), 1134 (w), 1085 (s), 1063 (s) 1038 (s), 1023 (m), 1011 (w), 997 (w), 989 (w), 897 (s), 880 (s), 831 (w), 782 (w), 754 (s), 738 (m), 695 (m), 673 (w), 618 (w), 589 (w). EI-MS: m/z 168 (90%) [NPh2]+. Anal. Found (calcd for C32H44N6Si2Ti): C, 62.2 (62.3); H, 7.2 (7.2); N, 13.7 (13.6). Ti(NNPh2)(N2NN′)(py) (13). To a stirred mixture of [Ti(NNPh2)Cl2(py)2]2 (0.30 g, 0.66 mmol) and Li2N2NN′ (0.23 g, 0.66 mmol) was added toluene (30 mL), all at -78 °C. The yellowgreen suspension was allowed to warm to room temperature and stirred for a further 90 min, resulting in a dark brown solution. Volatiles were removed under reduced pressure, and the brown solid was extracted into Et2O (2 × 5 mL) and filtered. Volatiles were removed under reduced pressure to yield a green-brown solid, which was washed with pentane (3 × 5 mL) cooled to -78 °C. The dark green solid product was dried in Vacuo. Yield: 0.12 g (31%). 1 H NMR (C6D6, 299.9 MHz, 293 K): δ 8.98 (1 H, d, 3J ) 6.5 Hz, 6-NC5H4), 7.75 (4 H, d, 3J ) 8.3 Hz, o-C6H5), 7.24 (4 H, app. t, app. 3J ) 8.3 and 7.6 Hz, m-C6H5), 6.87 (2 H, t, 3J ) 7.6 Hz, p-C6H5), 6.61 (1 H, app. t, app. 3J ) 7.6 and 7.7 Hz, 4-NC5H4), 6.12 (1 H, app. t, app. 3J ) 7.7 Hz, 5-NC5H4), 6.06 (1 H, d, 3J ) 7.6 Hz, 3-NC5H4), 3.66 (2 H, m, CH2NSiMe3 inner protons), 3.46 (2 H, m, CH2NSiMe3 outer protons), 3.01 (2 H, s, CH2NC5H4), 2.60 (2 H, m, CH2N inner protons), 2.08 (2 H, m, CH2N outer protons), 0.45 (18 H, s, SiMe3). 13C{1H} NMR (C6D6, 75.4 MHz, 293 K): δ 159.3 (2-NC5H5), 154.5 (6-NC5H5), 147.8 (i-C6H5), 140.0 (4-NC5H5), 129.5 (m-C6H5), 127.8, 123.5 (3-NC5H5 and 5-NC5H5), 122.0 (p-C6H5), 120.5 (o-C6H5), 57.3 (CH2NC5H5), 56.4 (NCH2), 49.2 (CH2NSiMe3), 3.2 (SiMe3). IR (NaCl plates, Nujol mull, cm-1): ν 1593 (m), 1583 (m), 1487 (s), 1312 (w), 1237 (m), 1168 (w), 1156 (w), 1086 (s), 1024 (m), 946 (m), 934 (m), 866 (m), 833 (s), 793 (m), 756 (w), 740 (m), 701 (w), 692 (w), 630 (w), 584 (w). EI-MS: m/z 168 (88%) [NPh2]+, 73 (88%) [SiMe3]. Anal. Found (calcd for C28H42N6Si2Ti): C, 59.2 (59.3); H, 7.4 (7.5); N, 14.7 (14.8).

Alternative Synthesis of Ti(NNPh2)(Me4taa) (14, NMR tube scale). To a solution of [Ti(NNPh2)Cl2(py)2]2 (12 mg, 13.1 µmol) in C6D6 (0.2 mL) was added a solution of Li2Me4taa (4.7 mg, 13.1 µmol) in C6D6 (0.2 mL). A color change of the initially cloudy solution to light brown was observed. Analysis by 1H NMR indicated that a reaction had occurred immediately and cleanly to give the known product, Ti(NNPh2)(Me4taa) (14).39 Ti(NNPh2)(O2NN′)(py) (15). To a stirred solution of [Ti(NNPh2)Cl2(py)2]2 (0.500 g, 1.09 mmol) in toluene (20 mL), cooled to -78 °C, was added a solution of Na2O2NN′ · 0.16(THF) (0.470 g, 1.09 mmol) in toluene (20 mL) at -78 °C. The solution was allowed to warm to room temperature and then stirred for a further 16 h. The volatiles were removed under reduced pressure to give an oily solid, which was extracted into toluene at -78 °C (2 × 20 mL) and filtered. Volatiles were removed under reduced pressure, and the product was washed with hexane (2 × 10 mL). The brown solid product was dried in Vacuo. Yield: 0.336 g (49%). 1 H NMR (C6D6, 299.9 MHz, 293 K): δ 9.21 (2 H, br, 2-C5H5N), 9.04 (1 H, d, 3J ) 6.3 Hz, 6-C5H4N), 8.31 (4 H, d, 3J ) 8.8 Hz, o-C6H5), 7.35 (4 H, app. t, app. 3J ) 7.5 and 8.8 Hz, m-C6H5), 6.94 (2 H, t, 3J ) 7.5 Hz, p-C6H5), 6.91 (2 H, d, 4J ) 2.1 Hz,

Organometallics, Vol. 27, No. 24, 2008 6493 4-C6H2Me2), 6.62 (1 H, broad, 4-C5H5N), 6.56 (2 H, d, 4J ) 2.1, 6-C6H2Me2), 6.38 (1 H, app. t, app. 3J ) 7.9 Hz, 4-C5H4N), 6.34 (2 H, br, 3-C5H5N), 6.06 (1 H, app. t, app. 3J ) 6.3 Hz, 5-C5H4N), 5.70 (1 H, d, 3J ) 7.9 Hz, 3-C5H4N), 3.51 (2 H, d, 2J ) 12.5 Hz, NCH2Ar distal to C5H4N), 3.20 (2 H, s, NCH2C5H4N), 2.59 (6 H, s, 3-C6H2Me2), 2.58 (2 H, d, NCH2Ar proximal to C5H4N), 2.25 (6 H, s, 5-C6H2Me2). 13C{1H} NMR (C6D6, 75.4 MHz, 293 K): δ 162.9 (2-C6H2Me2), 158.1 (2-C5H4N), 152.4 (6-C5H4N), 151.2(2C5H5N), 147.5 (i-C6H5), 138.0 (4-C5H4N), 132.2 (4-C5H2Me2), 129.5 (m-C6H5), 129.2 (6-C5H2Me2), 127.0 (5-C5H2Me2), 124.5 (3C5H5N), 124.3 (3-C5H2Me2), 123.2 (1-C5H2Me2), 122.0 (p-C6H5), 121.9 (5-C5H4N), 120.8 (3-C5H4N), 119.0 (o-C6H5), 63.4 (NCH2Ar), 58.1 (NCH2C5H4N), 21.1(5 C5H2Me2), 19.0 (3 C5H2Me2). IR (NaCl plates, Nujol mull, cm-1): ν 1604 (m), 1595 (s), 1583 (s), 1475 (s), 1444 (s), 1377 (m), 1318 (s), 1272 (s), 1214 (m), 1162 (m), 1102 (w), 1070 (w), 1053 (w), 1039 (m), 1023 (m), 992 (w), 960 (m), 856 (m), 822 (s), 749 (s), 729 (m), 694 (s), 648 (w), 632 (m), 587 (m), 548 (m). EI-MS: m/z 121 [OC6H2Me2]+ (99%), 105 [NNPh]+ (26%), 91 [NPh]+ (43%), 79 [NC5H5]+ (24%). Anal. Found (calcd for C41H41N5O2Ti): C, 71.9 (72.0); H, 6.0 (6.0); N, 10.2 (10.2). Crystal Structure Determinations of Ti(NNPh2)Cl2(py)3 (7), Ti(NNPh2)Cl2(py′)3 (8), Ti(NNPh2)(N2NC2,Me)(py) (2), Ti(NNPh2)(N2NC2,SiMe3)(py) (10), Ti(NNPh2)(N2NC3,Me)(py) (11), and Ti(NNPh2)(N2Npy)(py) (12). Crystal data collection and processing parameters are given in Table 6. Crystals were mounted on glass fibers using perfluoropolyether oil and cooled rapidly in a stream of cold N2 using an Oxford Cryosystems Cryostream unit. Diffraction data were measured using an EnrafNonius KappaCCD diffractometer. As appropriate, absorption and decay corrections were applied to the data and equivalent reflections merged.102 The structures were solved by direct methods (SIR92103), and further refinements and all other crystallographic calculations were performed using the CRYSTALS program suite.104 Other details of the structure solution and refinements are given in the Supporting Information (CIF data). A full listing of atomic coordinates, bond lengths and angles, and displacement parameters for all the structures has been deposited at the Cambridge Crystallographic Data Centre. See Notice to Authors, Issue No. 1. Computational Details. All the calculations have been performed with the Gaussian03 package105 at the B3PW91 level.106,107 The titanium atom was represented by the relativistic effective core potential (RECP) from the Stuttgart group (12 valence electrons) and its associated basis set,108 augmented by an f polarization function (R ) 0.869).109 The remaining atoms (C, H, N) were represented by a 6-31G(d,p) basis set.110 Full optimizations of geometry without any constraint were performed, followed by analytical computation of the Hessian matrix to confirm the nature of the located extremum as a minimum on the potential energy surface for the experimental complex 2. For the model systems Hyd_0, Hyd_90, Hyd_0_rot, Imide_90, and Imide_00, constraints were imposed in the geometry optimiza(102) Otwinowski, Z.; Minor, W. Processing of X-ray Diffraction Data Collected in Oscillation Mode; Academic Press: New York, 1997. (103) Altomare, A.; Cascarano, G.; Giacovazzo, G.; Guagliardi, A.; Burla, M. C.; Polidori, G.; Camalli, M. J. Appl. Crystallogr. 1994, 27, 435. (104) Betteridge, P. W.; Cooper, J. R.; Cooper, R. I.; Prout, K.; Watkin, D. J. J. Appl. Crystallogr. 2003, 36, 1487. (105) Frisch,M. J.; et al. Gaussian 03, ReVision C.02;Gaussian, Inc.: Wallingford, CT, 2004. (106) Becke, A. D. J. Chem. Phys. 1993, 98, 5648. (107) Perdew, J. P.; Wang, Y. Phys. ReV. B 1992, 45, 13244. (108) Andrae, D.; Haussermann, U.; Dolg, M.; Stoll, H.; Preuss, H. Theor. Chim. Acta 1990, 77, 123. (109) Ehlers, A. W.; Bohme, M.; Dapprich, S.; Gobbi, A.; Hollwarth, A.; Jonas, V.; Kohler, K. F.; Stegmann, R.; Veldkamp, A.; Frenking, G. Chem. Phys. Lett. 1993, 208, 111. (110) Hariharan, P. C.; Pople, J. A. Theor. Chim. Acta 1973, 28, 213.

6494 Organometallics, Vol. 27, No. 24, 2008

Selby et al.

Acknowledgment. We thank the EPSRC for support and Dr. A. R. Cowley for the X-ray structure of 8.

(7), Ti(NNPh2)Cl2(py′)3 (8), Ti(NNPh2)(N2NC2,Me)(py) (2), Ti(NNPh2)(N2NC2,SiMe3)(py) (10), Ti(NNPh2)(N2NC3,Me)(py) (11), Ti(NNPh2)(N2Npy)(py) (12), and Ti2(µ-η2η1NNMe2)2(O2NN′)2 · 0.5(C5H12) (16 · 0.5(C5H12)). Results and discussion for the attempted synthesis of other titanium hydrazides containing dianionic O2N2-donor ligands (16, eq S1, Figure S1, Tables S1-S2), comparison between experimental and calculated geometrical parameters for Ti(NNPh2)(N2NC2,Me)(py), 2 (Table S3), NPA charges for the model compounds Ti(NR)(NMe2)2(py)2 (Table S4), views of selected NLMOs for Hyd_0 (Figure S2), Cartesian coordinates and electronic energy for the optimized structures, complete list of authors for the reference of Gaussian03. This information is available free of charge via the Internet at http://pubs.acs.org.

Supporting Information Available: X-ray crystallographic data in CIF format for the structure determinations of Ti(NNPh2)Cl2(py)3

OM8007597

tion procedure. For all systems the amido groups NMe2 and the NβMe2 group (hydrazido models) were constrained to be planar, the pyridine rings were eclipsing the TidNR bond. The difference between the various systems lies in the orientation of the amido and/or the hydrazido NβMe2 groups, and these various constraints were imposed upon dihedral angles. In the label of the model systems “0” refers to amido groups lying in the equatorial plane, while “90” refers to amido groups perpendicular to this plane. The label “rot” in Hyd_0_rot implies that the NβMe2 group of the hydrazido is perpendicular to the equatorial plane. For all these systems NBO analyses were performed with NBO 5.0 interfaced with Gaussian.92